Vous êtes sur la page 1sur 279

THÈSE

En vue de l’obtention du

DOCTORAT DE L’UNIVERSITÉ DE TOULOUSE


Délivré par : l’Institut Supérieur de l’Aéronautique et de l’Espace (ISAE)

Présentée et soutenue le 01/12/2016 par :


Yann DENIEUL

Preliminary Design of Control Surfaces and Laws for Unconventional Aircraft


Configurations

JURY
Daniel ALAZARD Professeur, ISAE-SUPAERO Directeur de thèse

Joël Professeur associé, Co-directeur de thèse


BORDENEUVE-GUIBE ISAE-SUPAERO
Rogelio LOZANO Directeur de Recherche, UTC Rapporteur
de Compiègne
Florian HOLZAPFEL Professor, Technische Rapporteur
Universität München
Luc DUGARD Directeur de recherche CNRS, Examinateur
GIPSAS Grenoble
Andres MARCOS Associate Professor, University Examinateur
of Bristol
Clément TOUSSAINT Ingénieur de recherche à Examinateur
l’ONERA
Gilles TAQUIN Expert Qualités de vol, Airbus Examinateur

École doctorale et spécialité :


EDSYS : Automatique
Unité de Recherche :
ISAE-ONERA Commande des Systèmes et Dynamique du Vol -CSDV
Directeurs de Thèse :
Daniel ALAZARD et Joël BORDENEUVE-GUIBE
To my wife

i
Remerciements

“S’il n’y a pas de solution, c’est qu’il n’y a pas de problème.”


- Devise Shadok.

Cette devise, affichée sur mon bureau durant ces trois années de thèse, aurait pu être également
être la conclusion de ce travail sans le soutien et l’aide de nombreuses personnes que je tiens à
remercier ici. Commençons bien sûr par mes directeurs de thèse, et en premier lieu merci à toi
Daniel. En plus d’être à l’origine des idées qui ont mené à ce travail, ton encadrement toujours
avisé s’est avéré précieux dans les différentes étapes de cette thèse. Un immense merci à toi aussi
Joël, tu as su me guider avec sagesse et humanité dans les moments de doute et de remise en
question. Je tiens également à remercier Clément, notamment pour les longues discussions que
nous avons eues ensemble, et qui m’ont permis de prendre de la hauteur et du discernement
sur un sujet potentiellement piégeur par la diversité de problèmes qu’il soulève. Enfin je tiens à
exprimer mon immense gratitude envers Gilles, auprès de qui j’ai énormément appris tout au long
de ces années. Ta rigueur et ta pédagogie auront été des exemples pour moi ; je te suis également
reconnaissant de m’avoir fait confiance durant toutes ces années, comme quoi mon “bidule” de
H∞ a fini par aboutir !

Restons à Airbus, je garderai un souvenir ému de toute l’équipe des avant-projets que j’ai côtoyé
pendant trois ans et demi : un merci particulier à Marylène qui m’a initié à la beauté des ailes
volantes, Eric pour nos discussions sur comment changer le monde (et les outils d’Airbus),Thierry
qui n’a jamais ri à mes concepts d’avion les plus absurdes, Philippe qui a vainement tenté de me
faire partager son amour des turbopropulseurs, Serge avec qui les entretiens annuels ressemblent
plus à un colloque de théologie, mes voisines de box Sole, Claire et Jean-Pierre, Cécile, Thierry
S., Claire B., Mickaël, Christophe qui m’a appris l’art du pillage de pot de départ, Maud à qui
j’ai fait découvrir le monde merveilleux des AP, ainsi que bien sûr tous les djeuns : Romain,
Paul, Charly, JohnEtMarie, Charly, Jérôme et Jérôme, Sara, Augustin, Pierre-Yves, Elena, Peter,
Sebastian, et tous les stagiaires que j’ai côtoyés. J’ai également une pensée amicale également
pour les graisseux mais néanmoins collègues : Renaud que j’aurai pu charier avant qu’il ne devienne
mon chef, Mathieu et Mattieu, Laurent, Stéphane, Olivier, Guillaume, André, et tous ceux que
j’ai côtoyés et que je n’oublie pas. Plus largement à Airbus je remercie Pierre F. pour son soutien
technique et humain, ainsi que pour les ponts qu’il a contribué à bâtir entre nos deux services,
ainsi que Josep et Stéphane.

Revenons à Supaéro, je remercie Eric pour les discussions passionnées que nous avons eues
ensemble sur tout ce qui a trait à l’aéronautique, Stéphanie pour tes marques d’attention et ton
écoute, Marie pour ton aide au quotidien, ainsi que tout le personnel du DCAS et de l’ONERA.
Côté doctorants un bureau de folie s’est constitué au fil des années, et a largement contribué
à ma motivation journalière. Bastien pour avoir ramené au bureau le jeu le plus distrayant de
l’histoire de Supaéro ( et sans lequel j’aurais bouclé ma thèse en un an et demi au lieu de trois),
Vincent pour m’avoir prouvé qu’on pouvait camper dans un bureau, Emilien pour être à l’origine
du seul oonneeeee hhuuunnndddrreeeeddd aaannddd eeeiiiggghhttttyyyy que je verrai de ma vie.

iii
iv

J’ai également une pensée pour tous les doctorants de l’onera, Emmanuel, Alvaro, Martin, Jan et
tous les autres.

Bien sûr cette thèse aurait été très différente sans François, digne représentant de la Gauliste,
la horde des sans-manches, et toutes nos escalopes avec de belles salades.

Enfin comment ne pas terminer ces remerciements par celle qui aura à la fois été ma co-
directrice de thèse, première relectrice, référente aérodynamique (si même avec 400 mailles c’est
de la CFD !), ma femme : Amélo.
Table Of Contents

INTRODUCTION 3

I STATE-OF-THE-ART 9

1 State-of-the-Art 11

1.1 Design Challenges of Commercial Blended Wing-Body Aircraft . . . . . . 11

1.1.1 Overall Aircraft Design Aspects . . . . . . . . . . . . . . . . . 12

1.1.2 Stability and Control Challenges . . . . . . . . . . . . . . . . . 14

1.1.3 Flight Control System Sizing for Blended Wing-Body . . . . . . 19

1.2 Interactions of Aircraft Design and Control . . . . . . . . . . . . . . . . 24

1.2.1 Research in Control-Configured Vehicle . . . . . . . . . . . . . 24

1.2.2 Multidisciplinary Optimization Approach . . . . . . . . . . . . . 27

1.2.3 Control-Based Approach . . . . . . . . . . . . . . . . . . . . . 29

1.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

II METHODS AND TOOLS 33

2 Flight Dynamics Models 35

2.1 Flight Mechanics Equations . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.1.1 Coordinates Systems Definition . . . . . . . . . . . . . . . . . 35

2.1.2 Fundamental Kinematics Equation . . . . . . . . . . . . . . . . 39

2.1.3 Fundamental Mechanics Principle applied to the Airborne Aircraft 39

2.1.4 Longitudinal Model . . . . . . . . . . . . . . . . . . . . . . . . 42

2.1.5 Lateral Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.1.6 Concluding Remarks on Flight Dynamics Model . . . . . . . . . 49

2.2 Blended Wing-Body Geometrical Model . . . . . . . . . . . . . . . . . . 50

v
vi Table Of Contents

2.3 Description of Aircraft Flight Envelope . . . . . . . . . . . . . . . . . . . 52

2.4 Open Loop Modes Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 53

2.4.1 Longitudinal Modes . . . . . . . . . . . . . . . . . . . . . . . . 53

2.4.2 Lateral Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

2.5 Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3 Review on Control Laws Techniques and Application on an Academic Example 63

3.1 Presentation of the Academic Example: the Inverted Pendulum . . . . . 64

3.1.1 Open-Loop Dynamics . . . . . . . . . . . . . . . . . . . . . . . 65

3.1.2 Structure of the Stabilizing Control Law . . . . . . . . . . . . . 67

3.1.3 Actuator Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 68

3.2 Eigenvalues Assignment through Analytical Computation . . . . . . . . . 69

3.2.1 Analytical Computation Without Actuators Dynamics . . . . . 69

3.2.2 Inclusion of Actuators Dynamics . . . . . . . . . . . . . . . . . 70

3.3 Linear Quadratic Approach for Minimum Energy Control . . . . . . . . . 74

3.4 H∞ Formulation with Weighting on the Acceleration Sensitivity Function 75

3.4.1 Introduction to H∞ Control . . . . . . . . . . . . . . . . . . . 75

3.4.2 Acceleration Sensitivity Function (ASF) . . . . . . . . . . . . . 77

3.4.3 Full-Order H∞ Compensator . . . . . . . . . . . . . . . . . . 78

3.4.4 Structured H∞ Compensator . . . . . . . . . . . . . . . . . . 81

3.5 Co-design Approach on an Inverted Pendulum . . . . . . . . . . . . . . . 83

3.5.1 Optimal Actuator Bandwidth . . . . . . . . . . . . . . . . . . . 83

3.5.2 Minimal Stick Size . . . . . . . . . . . . . . . . . . . . . . . . 86

III SCIENTIFIC CONTRIBUTION 91

4 Elaboration of Design Procedures for Handling Qualities Sizing of an Unstable


Aircraft 93
Table Of Contents vii

4.1 Iterative Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

4.1.1 Aerodynamic Model . . . . . . . . . . . . . . . . . . . . . . . . 96

4.1.2 Handling Qualities . . . . . . . . . . . . . . . . . . . . . . . . 96

4.1.3 Actuators Pre-Sizing . . . . . . . . . . . . . . . . . . . . . . . 98

4.1.4 Systems Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . 102

4.2 Codesign Process for Control Surfaces Sizing . . . . . . . . . . . . . . . 102

4.3 Codesign Process for Actuators Bandwidth Sizing . . . . . . . . . . . . . 104

5 Combined Optimization of Stabilizing Longitudinal Control Law and Actuators


Bandwidth through H2 /H∞ Synthesis 107

5.1 Analytical Influence of Control Laws Gains and Actuators Bandwidth on


Deflection Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5.1.1 State-Space Representation of Different Models . . . . . . . . . 109

5.1.2 Full Linear Model . . . . . . . . . . . . . . . . . . . . . . . . . 112

5.1.3 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

5.2 Synthesis of Stabilizing Control Law based on the Acceleration Sensitivity


Function (ASF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

5.2.1 Description of Synthesis Scheme with Weighting on the Acceler-


ation Sensitivity Function . . . . . . . . . . . . . . . . . . . . . 115

5.2.2 Control Law Gains Structure . . . . . . . . . . . . . . . . . . 117

5.3 H∞ Synthesis with Optimization of Allocation Gains . . . . . . . . . . . 118

5.4 H2 /H∞ Synthesis with Optimization of Allocation Gains . . . . . . . . 121

5.5 Codesign of Control Law Gains and Actuators Bandwidth with Cost Func-
tion on Deflection Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

5.6 Evaluation of Maximum Aerodynamic Hinge Moments . . . . . . . . . . 130

5.7 Codesign of Control Law Gains and Actuators Bandwidth with Cost Func-
tion on Actuators Power Consumption . . . . . . . . . . . . . . . . . . . 130

5.8 Actuators Bandwidth Sizing on the Whole Flight Envelope . . . . . . . . 135

6 Three-Axes Aerodynamic Model for Handling Qualities Parameterized by Con-


viii Table Of Contents

trol Surfaces Size 139

6.1 Parameterized Geometrical Model of Blended Wing-Body Control Surfaces 140

6.2 Computation of Control Surfaces Aerodynamic Efficiency for Several Geo-


metrical Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6.2.1 Computation of Lift Coefficient . . . . . . . . . . . . . . . . . 142

6.2.2 Computation of Pitching Moment Coefficient . . . . . . . . . . 144

6.2.3 Computation of Rolling Moment Coefficient . . . . . . . . . . . 144

6.3 LFR Approximation of State-Space Representations . . . . . . . . . . . . 145

6.4 Computation of Maximum Hinge Moments for Evaluation of Power Con-


sumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

7 Co-Design of Control Surfaces and Laws under Maneuverability and Stability


Constraints 149

7.1 Development of a Three-Axes H∞ Model Reference Tracking Scheme . . 150

7.1.1 Model Reference Tracking for Longitudinal Synthesis . . . . . . 150

7.1.2 Reference Model Tracking for Three-Axes Synthesis . . . . . . 162

7.2 Translation of Handling Qualities Dynamic Constraints into H∞ Constraints169

7.3 Evaluation of Different Methods for Control Surfaces Codesign . . . . . . 174

7.3.1 Decoupled Optimization with LFR Parametrization of the Allo-


cation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

7.3.2 Codesign with LFR Parametrization of the Allocation . . . . . . 181

7.3.3 Codesign with Tunable Allocation . . . . . . . . . . . . . . . . 188

CONCLUSION AND PERSPECTIVES 199

APPENDIX 207

A Analytical Expression of State-Space Representations 207

A.1 Longitudinal Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . 207

A.2 Lateral Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208


Table Of Contents ix

RESUME ETENDU EN FRANCAIS 211

B Dimensionnement Conjoint de Surfaces de Contrôle et Lois de Commande


pour Configurations d’Avions Non-Conventionnelles 211

B.1 Description d’un processus de dimensionnement d’un avion instable au


stade avant-projets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

B.1.1 Processus itératif . . . . . . . . . . . . . . . . . . . . . . . . . 212

B.1.2 Processus couplé pour le dimensionnement des gouvernes . . . 214

B.2 Obtention d’un modèle aérodynamique paramétré . . . . . . . . . . . . . 216

B.2.1 Paramétrisation géométrique des gouvernes de bord de fuite . . 216

B.2.2 Calcul de modèles aérodynamiques pour des envergures de gou-


vernes discrétisées . . . . . . . . . . . . . . . . . . . . . . . . . 217

B.2.3 Approximation polynômiale et représentation LFR des efficacités


aérodynamiques de gouvernes . . . . . . . . . . . . . . . . . . 219

B.3 Codesign de gouvernes et lois de commandes sous contraintes qualités de vol221

B.3.1 Structure des lois de commande . . . . . . . . . . . . . . . . . 221

B.3.2 Modèle d’allocation . . . . . . . . . . . . . . . . . . . . . . . . 221

B.3.3 Synthèse simultanée de lois 3 axes . . . . . . . . . . . . . . . . 222

B.3.4 Définition du problème d’optimisation . . . . . . . . . . . . . . 226

B.4 Résultats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

B.4.1 Codesign sur un seul point de vol . . . . . . . . . . . . . . . . 229

B.4.2 Résultats sur l’ensemble du domaine de vol . . . . . . . . . . . 229

BIBLIOGRAPHY 232
List of Figures

1.1 Flying wings prototypes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.2 Conceptual view on wetted area gains of a BWB vs conventional aircraft. Source:
[Liebeck, 2004] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.3 Boeing BWB concept. Source: [Liebeck, 2004] . . . . . . . . . . . . . . . . . . 14

1.4 Structural concepts for BWB pressurized shells. Source: [Liebeck, 2004]. . . . . 15

1.5 Airbus BWB concept. Source: [Saucez, 2013] . . . . . . . . . . . . . . . . . . . 16

1.6 Unmanned tailless aircraft. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.7 Geometry of baseline and CCV tanker. Source: [Walker, 1976]. . . . . . . . . . . 26

1.8 Comparison of optimized designs, without and with flight dynamics and control
constraints. Source [Perez et al., 2006] . . . . . . . . . . . . . . . . . . . . . . 28

2.1 Earth and Aircraft Body Reference Systems. . . . . . . . . . . . . . . . . . . . . 37

2.2 Aerodynamic and Aircraft Body Reference System. . . . . . . . . . . . . . . . . 38

2.3 Aerodynamic and Earth Reference Systems. . . . . . . . . . . . . . . . . . . . . 39

2.4 Planform view of the Airbus Blended Wing-Body. . . . . . . . . . . . . . . . . . 50

2.5 Illustration of significant hinge of elevon 2. . . . . . . . . . . . . . . . . . . . . . 52

2.6 Mach-Altitude flight envelope. . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

2.7 Comparison of Longitudinal Poles with Reduced State-Space Representations [α, q]


and [V, θ]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

2.8 Position of aerodynamic center XF with respect to Center of Gravity (CG) in


high and low speed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

2.9 Evolution of Longitudinal Poles with Speed. . . . . . . . . . . . . . . . . . . . . 57

2.10 Evolution of Lateral Poles with Speed. . . . . . . . . . . . . . . . . . . . . . . . 59

2.11 σz as function of altitude and probability of exceedance, according to [Chalk,


1969] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

3.1 Inverted Pendulum. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

xi
xii List of Figures

3.2 Block-diagram of pendulum dynamics and stabilizing control law. . . . . . . . . 66

3.3 Open-loop poles location of inverted pendulum and effect of a position feedback
with gain k1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

3.4 Temporal and frequency responses of a first-order actuator with bandwidth ωact =
5Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

3.5 Impulse Responses of controlled inverted pendulum with different control tech-
niques and ωact = 5Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.6 Poles of controlled inverted pendulum with different control techniques and ωact =
5Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.7 Impulse Responses of controlled inverted pendulum with different control tech-
niques and ωact = 1.75Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

3.8 Poles of controlled inverted pendulum with different control techniques and ωact =
1.75Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

3.9 Standard form for H∞ control problem. . . . . . . . . . . . . . . . . . . . . . . 75

3.10 Augmented Standard Form with Weighting on Acceleration Sensitivity Function


(ASF). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

3.11 Bode magnitude diagrams of inverse of frequency template 1/W1 , and Acceleration
Sensitivity Function (ASF) with full-order and structured H∞ compensators, for
ωact = 5Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

3.12 Bode magnitude diagrams of inverse of frequency template 1/W1 , and ASF with
full-order and structured H∞ compensators, for ωact = 1.75Hz. . . . . . . . . . 79

3.13 Block-diagram for co-design of control law gains k1 and k2 , and actuator
bandwidth ωact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

opt 1.1
3.14 kTw→z (P, k1 , k2 , ωact )k∞ , in blue, and constraint W 1
in yellow. Constraint is
fulfilled if blue curve is below yellow zone. . . . . . . . . . . . . . . . . . . . . . 85

3.15 Illustration of co-design problem. . . . . . . . . . . . . . . . . . . . . . . . . . . 86

3.16 Block-diagram for co-design of control law gains k1 , k2 and minimum stick size
lp . Tunable blocks are colored in orange. . . . . . . . . . . . . . . . . . . . . . 87

3.17 Root Mean-Square (RMS) value of angular position θp for disturbance Wp


with RMS value of 5NM (in red), and γ∞ struct after structured synthesis (in blue),

for varying values of stick length lp . . . . . . . . . . . . . . . . . . . . . . . . . 88


List of Figures xiii

4.1 Iterative process. The time scale is increasing from top to bottom. Each box
represents a discipline of the process. Inputs for each discipline are represented
by arrows coming to the top of each box; outputs of each discipline are arrows
coming from the bottom of each box. Constraints for each discipline are coming
to the side of each box. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

4.2 Schematic view of control surface actuator installation. . . . . . . . . . . . . . . 98

4.3 Schematic actuator capabilities curve: deflection rate δm versus hinge moment
HM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

4.4 Codesign process for control surfaces sizing. The main difference between this
figure and the iterative process description is that Maximum Deflections and De-
flection Rates are now constraints of the Closed-Loop Handling Qualities Process
instead of outputs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

4.5 Codesign process for actuators bandwidth sizing. . . . . . . . . . . . . . . . . . 105

5.1 Block-diagram of stabilizing control law computed with ASF scheme. Fixed
parameters are in green, and tunable parameters are in red. . . . . . . . . . . . . 116

5.2 Temporal simulation in response to an initial condition α0 on angle of attack, with


control law computed through H∞ synthesis. . . . . . . . . . . . . . . . . . . . 119

5.3 Control allocation gains obtained with H∞ synthesis, and control surfaces pitch
effectiveness. mδmi is the i-th elevon pitch evectiveness defined in equation (5.34). 120

5.4 Bode Magnitude response of the ASF after H∞ optimization. . . . . . . . . . . 120

5.5 Temporal simulation in response to an initial condition α0 on angle of attack, with


control law computed through H2 /H∞ synthesis. . . . . . . . . . . . . . . . . . 122

5.6 Control allocation gains obtained with H2 /H∞ synthesis, and control surfaces
pitch effectiveness. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

5.7 Bode Magnitude response of the ASF after H2 /H∞ optimization. . . . . . . . 123

5.8 Block-diagram of stabilizing control law and actuators bandwidth optimization


computed with ASF scheme. Fixed parameters are in green, and tunable param-
eters are in orange. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

5.9 Temporal simulation in response to an initial condition α0 on angle of attack, with


control law and actuators bandwidth computed through codesign with objective
set on kTw→Uad k2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

5.10 Optimized actuators bandwidth in red, plotted together with normalized hinge
moments in blue, and normalized elevons pitch effectiveness, in green. . . . . . . 128
xiv List of Figures

5.11 Bode Magnitude response of the ASF after H2 /H∞ codesign on actuators
bandwidth. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

5.12 Maximum hinge moment, computed for δmi = 30˚, as a function of flight point
for each elevon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

5.13 Temporal simulation in response to an initial condition α0 on angle of attack, with


control law and actuators bandwidth computed through codesign with objective
set on kTw→WHM ×Uad k2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

5.14 Optimized actuators bandwidth in red, plotted together with normalized hinge
moments in blue, and normalized elevons pitch effectiveness, in green. Values are
obtained after an optimization with a minimization objective on kTw→WHM ×Uad k2 .133

5.15 Bode Magnitude response of the ASF after H2 /H∞ codesign on actuators
bandwidth with weighting on the actuators power consumption. . . . . . . . . . 133

5.16 Actuators optimal bandwidth after codesign with hinge moments weighting, for all
flight points. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

5.17 Maximum hinge moment, computed for δmi = 30˚, as a function of flight point
for each elevon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.1 Elevons size for different values of parameter η. . . . . . . . . . . . . . . . . . . 140

6.2 Process for obtaining LFR approximation of state-space representations as a func-


tion of η parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6.3 Comparison of pitch gradient coefficients CLδmi from AVL outputs with reference
aircraft aerodynamic data (left) and with calibrated data (right). . . . . . . . . . 142

6.4 Panels distribution for the Blended Wing-Body (BWB) lifting surfaces. . . . . . 143

6.5 Linear Fractional Representation. . . . . . . . . . . . . . . . . . . . . . . . . . . 145

6.6 LFR polynomial approximations of elevons pitch and roll gradients. . . . . . . . . 146

7.1 Longitudinal model reference tracking scheme, with longitudinal gains structure
and error computed on the actual nz signal: z1 = nref
z − nz . . . . . . . . . . . 152

7.2 Bode diagram and poles representation on complex plane, for longitudinal reference
model and closed-loop resulting from minimization of Tnzc →z1 ∞ . . . . . . . . 153

7.3 Temporal responses for a step input δnzc = 1g, with control law computed by
minimization of Tnzc →z1 ∞ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
List of Figures xv

7.4 Bode diagram and poles representation on complex plane, for longitudinal refer-
ence model and closed-loop resulting from minimization of Tnzc →z1 ∞ with a
constraint set on poles location. . . . . . . . . . . . . . . . . . . . . . . . . . . 156

7.5 Temporal responses for a step input δnzc = 1g, with control law computed by
minimization of Tnzc →z1 ∞ and adding a constraint on poles location. . . . . . 157

7.6 Longitudinal model reference tracking scheme, with longitudinal gains structure
and error computed on approximated ñz signal: z̃1 = nref
z − ñz . . . . . . . . . 159

7.7 Bode diagram and poles representation on complex plane, for longitudinal refer-
ence model and closed-loop resulting from minimization of Tnzc →z̃1 ∞ with a
constraint set on poles location. . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7.8 Temporal responses for a step input δnzc = 1g, with control law computed by
minimization of Tnzc →z̃1 ∞ and adding a constraint on poles location. . . . . . 161

7.9 Three-axes Reference Model Tracking (RMT) scheme. Tunable blocks are dis-
played in orange. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

7.10 Linear Fractional Representation (LFR) polynomial approximations of control


allocation matrix Kalloc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

7.11 Comparison of reference models and temporal responses on nz , φ and β to a


unit step input in nzc , φc and βc respectively. . . . . . . . . . . . . . . . . . 166

7.12 Bode diagrams for three-axes reference models and closed-loop resulting from min-
imization of T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } and with a constraint on closed-loop poles

location. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

7.13 Temporal responses to inputs of δnzc = 1g at t = 0s, φc = 30˚at t = 5s, and βc =


5˚at t = 12s, with gains computed by a minimization of T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 }

under constraint of closed-loop poles location. . . . . . . . . . . . . . . . . . . . 168

7.14 H∞ constraints from φc to δmi normalized. . . . . . . . . . . . . . . . . . . 171

7.15 Control surfaces deflections to a sinusoidal input of φc at frequency 0.79 rad.s−1 ,


for η opt = 0.4165 computed so that ∆δm1i max Tφc →u φmax ≤ 1, with ∆δmi max =

10˚ and φmax = 30˚. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

7.16 Control surfaces deflections to a step input of φc at t = 5s, for η opt = 0.4165
computed so that ∆δm1i max Tφc →u φmax ≤ 1, with ∆δmi max = 10˚and φmax =

30˚. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

7.17 Temporal response of control surfaces to severe longitudinal turbulence, with η opt =
0.3966 computed so that δmi2max Tew →u ≤ 1, with δmi max = 5˚. . . . . . . 173

7.18 Scheme for decoupled optimization. Tunable blocks are displayed in orange. . . . 175
xvi List of Figures

7.19 Bar of constraints values after decoupled optimization, η opt = 0.7415. . . . . . . 177

7.20 Comparison of temporal responses for η = 1 (dotted lines) and η opt = 0.7415
(solid lines) computed through decoupled optimization. . . . . . . . . . . . . . . 179

7.21 Amount of times each constraint is limiting, on the whole flight envelope (156
flight points), after decoupled optimization. A constraint is considered limiting if
its normalized value is above 0.99. The sum of limiting constraints may exceed
the number of flight points, for several constraints may be active for a single flight
point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

7.22 η opt as a function of flight point, computed with decoupled optimization, together
with associated limiting constraints. . . . . . . . . . . . . . . . . . . . . . . . . 181

7.23 Codesign with parameterized control allocation. . . . . . . . . . . . . . . . . . . 182

7.24 Comparison of temporal responses for η = 1 (dotted lines) and η opt = 0.4161 (solid
lines) computed through codesign with LFR parametrization of the allocation. . 184

7.25 Bar of constraints values after codesign with parameterized allocation, η opt = 0.4161.185

7.26 η opt as a function of flight point, computed with codesign with parameterized
allocation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

7.27 Amount of times each constraint is limiting, on the whole flight envelope (156
flight points), with parameterized codesign. A constraint is considered limiting if
its normalized value is above 0.99. The sum of limiting constraints may exceed
the number of flight points, for several constraints may be active for a single flight
point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

7.28 Codesign with tunable control allocation. . . . . . . . . . . . . . . . . . . . . . . 188

7.29 Comparison of temporal responses for η = 1 (dotted lines) and η opt = 0.3885
(solid lines) computed through codesign with tunable control allocation. . . . . . 190

7.30 Bar of constraints values after codesign with tunable allocation, η opt = 0.3885. . 192

7.31 η opt as a function of flight point, computed with codesign with tunable allocation.193

7.32 Amount of times each constraint is limiting, on the whole flight envelope (156
flight points), with tunable allocation codesign. A constraint is considered limiting
if its normalized value is above 0.99. The sum of limiting constraints may exceed
the number of flight points, for several constraints may be active for a single flight
point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

7.33 Comparison of η opt with three optimization methods, for all flight points. . . . 195
List of Figures xvii

B.1 Processus itératif. L’échelle temporelle est représentée croissante de haut en bas.
Chaque boîte représente une discipline du processus. Les entrées de chaque disci-
plines sont représentées par des flèches arrivant par le haut de chaque boîte ; les
résultats de chaque discipline sont les flèches provenant du bas de chaque boîte.
Quant aux contraintes spécifiques à chaque discipline, proviennent du côté. . . . 213

B.2 Processus couplé pour le dimensionnement des gouvernes. La différence principale


entre cette figure et celle du processus dit “itératif” réside dans le rôle des déflec-
tions et vitesses de débattement maximales : elles sont désormais vues comme des
contraintes de la discipline Qualités de Vol Boucle Fermée, et non plus comme
des résultats. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215

B.3 Taille des gouvernes externes pour diférentes valeurs du paramètre η. . . . . . . 216

B.4 Processus d’obtention de représentation d’état paramétrées sous formalisme LFR,


avec comme paramètre variant η. . . . . . . . . . . . . . . . . . . . . . . . . . . 219

B.5 Comparaison des gradients de portance CLδmi en sortie d’AVL avec les données de
référence (gauche) et les données calibrées (droite). . . . . . . . . . . . . . . . . 220

B.6 Approximations polynômiales LFR des gradients de tangage et roulis. . . . . . . 220

B.7 Schéma de principe du codesign avec allocation paramétrée. Les blocs tunables
sont présentés en orange. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223

B.8 Diagrammes de Bode en amplitude selon les 3 axes des modèles de référence
et boucle fermée résultant de la minimisation de T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } avec

contrainte sur le lieu des pôles. . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

B.9 Réponses temporelles pour η = 1 (pointillés) et η opt = 0.3885 (traits pleins). . . 227

B.10 Diagramme en barres des contraintes normalisées après optimisation, η opt = 0.3885.230

B.11 η opt en fonction du cas de vol, calculé via le codesign avec allocation tunable. . 231
List of Tables

2.1 Top-Level Aircraft Requirement for Airbus Long-Range BWB . . . . . . . . . . . 50

2.2 Geometry data of the configuration. . . . . . . . . . . . . . . . . . . . . . . . . 51

2.3 Control Surfaces Attributes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

3.1 Pendulum numerical values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3.2 Gains k1 and k2 computed with different control techniques, for two actuator
bandwidth. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

5.1 Gains values for different control synthesis techniques. All values are shown mul-
tiplied by max Kalloc for each synthesis, in order to make fair comparisons. . . . 121

5.2 Gains values computed by codesign optimization, without and with weighing by
hinge moments. All values are shown multiplied by max Kalloc for each synthesis,
in order to make fair comparisons. . . . . . . . . . . . . . . . . . . . . . . . . . 134

7.1 Values for lateral and directional reference models. . . . . . . . . . . . . . . . . 163

7.2 Constraints values after decoupled optimization. . . . . . . . . . . . . . . . . . . 176

B.1 Valeurs pour le modèle de référence. . . . . . . . . . . . . . . . . . . . . . . . . 224

xix
Glossary

Fthruste Propulsive Forces at Equilibrium. 45

Fthrust Propulsive Forces. 41, 43


G
Mthrust Norm of Propulsive Moments applied on Aircraft Center of Gravity. 42

Faero Aerodynamic Forces. 40, 41


G
Maero Aerodynamic Moments applied on the Aircraft Center of Gravity. 40, 42
G
Mthrust Propulsive Moments applied on Aircraft Center of Gravity. 40, 42

Vg Aircraft Ground Speed. 40, 43

Vw Wind Speed. 40

ΩG Aircraft Rotation Vector around its Center of Gravity. 40

Aa Aircraft State Matrix. 46, 109, 112, 113

Au Actuator State Matrix. 110, 112–114

Aw Wind State Matrix. 112, 113

A State Matrix. 30, 46, 55, 145, 220

Ba Aircraft Control Matrix. 46, 109, 112, 113

Bu Actuator Control Matrix. 110, 112–114

Bw Wind Control Matrix. 112, 113

Baw Wind Influence on Aircraft States. 109, 112, 113

B Control Matrix. 30, 46, 124, 125, 145, 220

Ca Aircraft Output Matrix. 47, 109, 112–114

Cu Actuator Output Matrix. 110, 112, 113

Cw Wind Output Matrix. 112, 113

CHMi I-th Control Surface Aerodynamic Hinge Moment Coefficient. 146, 147

ClG Aerodynamic Rolling Moment Coefficient on the Center of Gravity. 42, 48

Clp Aerodynamic Rolling Moment Gradient with respect to Roll Rate. 48, 49, 58, 59

Clr Aerodynamic Rolling Moment Gradient with respect to Yaw Rate. 48, 49, 208

xxi
xxii Glossary

Clβ Aerodynamic Rolling Moment Gradient with respect to Sideslip. 48, 49

Clδmi Aerodynamic Rolling Moment Gradient with respect to i − th Elevon Deflection. 48, 49,
144, 145, 218

Clδn Aerodynamic Rolling Moment Gradient with respect to Rudder Deflection. 208

Cl Aircraft Rolling Moment Coefficient. 144, 145, 218

CmG
e Aerodynamic Pitching Moment at Equilibrium. 45

CmG
δmi i-th Elevon Pitch Gradient. 44, 45, 144, 207, 218

CmG Aerodynamic Pitching Moment Coefficient on the Center of Gravity. 42–44

Cm0 Zero angle-of-attack pitching moment. 3, 17–19, 44

CmG
q Pitch Rate Damping. 44, 45, 55, 56, 66, 207

CmG
α Angle of Attack gradient of Aerodynamic Pitching Moment. 45, 55, 56, 207

Cm Aircraft Pitching Moment Coefficient. 144, 217, 218

CnG Aerodynamic Yawing Moment Coefficient on the Center of Gravity. 42, 48

Cnp Aerodynamic Yawing Moment Gradient with respect to Roll Rate. 48, 49, 208

Cnr Aerodynamic Yawing Moment Gradient with respect to Yaw Rate. 48, 49, 60, 208

Cnβ Aerodynamic Yawing Moment Gradient with respect to Sideslip. 48, 49, 208

Cnδn Aerodynamic Yawing Moment Gradient with respect to Rudder Deflection. 48, 49

Cx0 Form Drag. 44

Cxe Drag Coefficient at Equilibrium. 45

Cx Drag Coefficient. 41, 43–45, 207

Cya Lateral Aerodynamic Force Coefficient Projected on Aerodynamic Frame.. 41

Cyp Aerodynamic Lateral Force Gradient with respect to Roll Rate. 48, 49

Cyr Aerodynamic Lateral Force Gradient with respect to Yaw Rate. 48, 49

Cyβ Aerodynamic Lateral Force Gradient with respect to Sideslip. 48, 49, 60, 208

Cyδn Aerodynamic Lateral Force Gradient with respect to Rudder Deflection. 48, 49

Cy Lateral Aerodynamic Force Coefficient Projected on Aircraft Frame.. 41, 48

Cz buffet Buffeting Lift Coefficient. 52

Cze Lift Coefficient at Equilibrium. 45


Glossary xxiii

Czα Angle of Attack Lift Gradient. 44, 45, 55, 66, 207

Czδmi i-th Elevon Lift Gradient. 44, 45, 66, 142–144, 147, 207, 217, 218, 220

Czmax Maximum Lift Gradient. 25, 52

Czq Pitch Rate Lift Gradient. 44, 45, 55, 56, 66, 207

Cz Aircraft Lift Coefficient. 18, 41, 43, 44, 142, 143, 147, 217

C Output Matrix. 46

Da Aircraft Direct Feedthrough Matrix. 47, 109, 112, 113

Du Actuator Direct Feedthrough Matrix. 110

Daw Wind Influence on Aircraft Outputs. 109, 110, 112–114

D Direct Feedthrough Matrix. 46

HMi i-th Control Surface Aerodynamic Hinge Moment. 134, 146, 147

HM Control Surface Aerodynamic Hinge Moment. xiii, 98–101, 128, 130, 133, 134, 137, 146

H Altitude. 52, 118, 153, 176, 183, 185, 191, 225, 229

Ki Gain on Load Factor Integral Error. 152, 155, 158, 160, 164

Kq Pitch Rate Feedback. 111, 112, 114, 117, 118, 120, 121, 123, 124, 129, 131, 134, 152, 155,
158, 160, 164

Kβc Feedforward Gain of Commanded Sideslip to Equivalent Rudder. 164

Kφc Feedforward Gain of Commanded Bank Angle to Equivalent Aileron. 164

Kθ Pitch Attitude Feedback. 111, 112, 114, 117, 118, 121, 124, 129, 131, 134

Kalloc Allocation Matrix. xv, xix, 111, 117, 118, 120, 121, 123–125, 129, 134, 164, 165, 189,
191, 222

Kipβ Feedback gain of Integral Error on Sideslip to Equivalent Aileron. 164

Kirβ Feedback Gain of Integral Error on Sideslip to Equivalent Rudder. 164

Knzc Load Factor Feedforward Gain. 152, 155, 158, 160, 164

Knz Load Factor Feedback. 111, 112, 114, 117, 118, 121, 124, 129, 131, 134, 152, 155, 158,
160, 164

Kpβ Feedback gain of Sideslip to Equivalent Aileron. 164

Kpφ Feedback gain of Bank Angle to Equivalent Aileron. 164

Kpp Feedback gain of Roll Rate to Equivalent Aileron. 164


xxiv Glossary

Kpr Feedback gain of Yaw Rate to Equivalent Aileron. 164

Krβ Feedback Gain of Sideslip to Equivalent Rudder. 164

Krφ Feedback Gain of Bank Angle to Equivalent Rudder. 164

Krp Feedback Gain of Roll Rate to Equivalent Rudder. 164

Krr Feedback Gain of Yaw Rate to Equivalent Rudder. 164

Kwact Vector of Tunable Bandwidth. 129, 131

K Gains Matrix. 74, 111, 113, 114, 163–165, 171, 174, 175, 181, 182, 188, 189, 224, 228

Lz Scale Length of Vertical Turbulence. 60–62, 112

M Mach Number. 52, 118, 185, 191, 225, 229

PCP Actuator Corner Power. 99–101

Pact Actuator Power Consumption. 99, 101

R0 Earth Reference System. 35, 36, 38, 40, 42, 43, 60

Ra Aerodynamic Reference System. 36–38, 41–43

Rb Aircraft Body Reference System. 36, 37, 41, 42, 47

Rxx x−wise Inertia Radius of the Aircraft. 41, 48, 49, 58, 208

Rxz xz−wise Cross-Inertia Radius of the Aircraft. 41, 47

Ryy y−wise Inertia Radius of the Aircraft. 41, 43, 45, 48, 207

Rzz z−wise Inertia Radius of the Aircraft. 41, 48, 49, 208

Swet Wetted area. 12

S Reference Area. 41–43, 45, 48, 49, 51, 56, 58, 98, 146, 207

Tact Actuator Transfer Function. 68, 69

T Temperature. 52

Ua Actuated Deflections Vector. 46, 109, 110, 113, 153

Uw Wind Output in Earth Reference Frame. 109, 112

Uad Actuated Deflection Rates Vector. xiii, xiv, 113, 114, 121, 124, 127, 129–134, 137

U Controls Vector. 46

Vair Airspeed. 21, 40–44, 46, 47, 52–54, 57, 58, 60, 61, 98, 109, 146

Ve Airspeed at Equilibrium. 45, 46, 48, 49, 57, 58, 109, 112, 159, 207, 208
Glossary xxv

Wp Pendulum Torque Perturbation. xii, 64, 65, 70, 87–90

WHM Weighting Matrix of Hinge Moments. xiv, 131–134

Wdes Desired Closed-Loop Model. 68, 80, 82, 88

Xa Aircraft State Vector. 46, 47, 109, 112, 113

Xu Actuator State Vector. 110, 112, 113

Xw Wind State Vector. 112, 113

XCG Center of Gravity x−wise location. 17, 18, 44, 55, 56, 144, 218

XFi x−wise location of i-th Control Surface Aerodynamic Center. 144, 218

XFq Maneuver Point x−wise location. 56

XF Aerodynamic Neutral Point x−wise location. xi, 17–19, 44, 55–57, 96

Xcp Center of Pressure x−wise location. 17, 18

X State Vector. 46, 74

Ya Aircraft Output Vector. 47, 109, 111–113

Y Output Vector. 46

Ψ Azimuth Angle. 37

α0 Zero Lift Angle of Attack. 44

αe Angle of Attack at Equilibrium. 45, 46, 48, 111

α Angle of Attack. 17, 18, 21, 29, 30, 38, 43–47, 49, 53, 54, 66, 109, 111, 115–117, 119, 122,
127, 132, 142–144, 158, 159, 217, 218

σ̄ Maximal Singular Value. 76

βc Commanded Sideslip. xv, xvii, 162–168, 171, 174, 175, 181–183, 188, 189, 221, 223–226, 228

βref Reference Model for Sideslip. 163, 167, 226

β Sideslip. xv, 37, 42, 48, 49, 59, 163, 164, 166–168, 179, 184, 190, 223, 226, 227

Fthrust Propulsive Forces Vector.. 40, 41

J G Aircraft Inertia Matrix expressed on its Center of Gravity. 40, 41

ΩRb /R0 Rotation Vector of Aircraft Body Frame with respect to Reference Frame. 36, 42

g Gravity Vector. 40

n Load Factor Vector. 40, 41


xxvi Glossary

x0 x−wise orthonormal vector of the Earth Reference Frame. 36–38

xa x−wise orthonormal vector of the Aerodynamic Reference Frame. 37, 38, 40, 43, 45

xb x−wise orthonormal vector of the Aircraft Reference Frame. 36–38, 41, 42, 47, 48

y0 y−wise orthonormal vector of the Earth Reference Frame. 36–38, 43

ya y−wise orthonormal vector of the Aerodynamic Reference Frame. 37, 38, 43, 45

yb y−wise orthonormal vector of the Aircraft Reference Frame. 36, 37, 42, 43, 47, 48

z0 z−wise orthonormal vector of the Earth Reference Frame. 36, 40, 42

za z−wise orthonormal vector of the Aerodynamic Reference Frame. 37, 43, 45

zb z−wise orthonormal vector of the Aircraft Reference Frame. 36, 41, 42, 47, 48

χ Aerodynamic Azimuth. 38

δαw Angle-of-Attack Increase due to Vertical Turbulence. 109

δlequi Equivalent Aileron Deflection. 164, 221, 222

δmi i-th Elevon Deflection.. xiv, xv, 44–46, 48, 49, 58, 109, 115, 117, 125, 131, 137, 142–147,
153, 170–174, 176, 177, 180, 182, 185, 187, 189, 192, 194, 217, 218, 228, 230

δmequi Equivalent Elevator Deflection. 117, 125, 152, 164, 221, 222

δmie i-th Elevon Trim Deflection. 46, 170, 228

δm Elevon Deflection.. xiii, 98–101, 119, 122, 127, 132, 135, 154, 157, 161, 168, 171–173, 177,
179, 180, 184, 185, 187, 190, 192, 194, 227, 230

δnequi Equivalent Rudder Deflection. 164, 221, 222

δn Rudder Deflection. 48, 49, 154, 157, 161, 168, 173, 179, 184, 190, 227

δxe Thrust Control at Equilibrium. 46

δx Thrust Control. 46
˙ i i-th Elevon Deflection Rate.. 170, 176, 177, 180, 182, 185, 187, 189, 192, 194, 228, 230
δm
˙ Elevon Deflection Rate.. 99, 100, 114, 119, 122, 127, 129, 132, 134, 154, 157, 161, 168,
δm
179, 184, 190, 227
˙ Rudder Deflection Rate. 154, 157, 161, 168, 179, 184, 190, 227
δn

η opt Optimized Value for Relative Span of Trailing-Edge Control Surfaces. xv–xvii, 171–174,
176–187, 189–195, 227, 229–231

η Total Relative Span of Trailing-Edge Control Surfaces. xvi, xvii, 140–142, 145, 146, 150, 153,
164, 165, 169–171, 174–176, 178, 179, 181–184, 186, 188–190, 216, 217, 219–228
Glossary xxvii

γ air Adiabatic Index. 52

γe Flight Path Angle at Equilibrium. 45, 46

γ∞ Optimal H∞ Norm. 78, 80, 84, 118, 121, 124, 129, 131, 155, 158, 160, 166, 171, 174–176,
181–183, 188, 189, 191, 225, 226, 228

γ Flight Path Angle. 38, 43, 45, 46, 53, 54, 66

C Field of Complex Numbers. 156, 158, 160, 165, 171, 174, 175, 181, 182, 188, 189, 224

R Field of Real Numbers. 76, 82, 84, 89, 118, 124, 129, 131

µ Aerodynamic Bank Angle. 38, 42

ω0 Desired Closed-Loop Frequency. 68–71, 77, 78, 80, 118, 151

ωact Actuator Bandwidth. xii, 68–73, 75, 78–80, 82–84, 89, 90, 110, 114, 128, 129, 133, 153

ωdr Dutch Roll Mode Desired Pulsation. 163

φc Commanded Bank Angle. xv, xvii, 162–168, 170–177, 180–183, 185, 187–189, 192, 194, 221,
223–226, 228, 230

φref Reference Model for Bank Angle. 162, 167, 226

φ Bank Angle. xv, 37, 42, 47–49, 59, 163, 164, 166–168, 179, 184, 190, 223, 226, 227

ρ Air Volumetric Mass. 41–43, 45, 48, 49, 56–58, 98, 146, 207, 208

σz Root-mean-square Intensity of Vertical Turbulence. xi, 60–62, 112, 174

τRP Specification for Pure Roll Mode Pulsation. 162

τSP Specification for Spiral Mode Pulsation. 162

θc Pendulum Commanded Angle with respect to Vertical Axis. 67, 69, 70

θp Pendulum Angle with respect to Vertical Axis. xii, 64, 65, 67, 69, 70, 72, 73, 78, 86–90

θ Pitch Attitude. 21, 37, 43, 46–49, 54, 66, 111, 115–117, 119, 122, 127, 132, 150

ñz Approximated Load Factor Based on Angle-of-Attack. xv, 158, 159, 163

z̃1 Error on Longitudinal Reference Model with Approximated Load Factor. xv, xvii, 159–161,
163, 165, 167, 168, 171, 174, 175, 181–183, 188, 189, 224, 226, 228

ξ0 Desired Closed-Loop Damping. 68–71, 77, 78, 80, 118, 151

ξact Actuator Damping. 153

ξdr Dutch Roll Desired Damping. 163

b Aircraft wing span. 12, 51


xxviii Glossary

ew White Noise with Unit Power for Dryden Turbulence Filter Input. xv, 60, 61, 112, 113, 170,
173, 174, 176, 177, 180, 182, 185, 187, 189, 192, 194, 228, 230

gBest Value of Optimal Constraints at the Optimum. 158, 160, 166, 225, 229

hopt
2 Optimal H2 Norm. 124, 126

k1 Angular Position Feedback Gain for Pendulum Control. xii, xix, 67, 69–71, 81–85, 87–89

k2 Angular Speed Feedback Gain for Pendulum Control. xii, xix, 67, 69–71, 81–85, 87–89

ki Induced Drag Factor. 44, 45, 207

lp Pendulum length. xii, 64–67, 69–71, 74, 75, 87–90

l Reference Length. 17, 18, 42–45, 48, 49, 55, 56, 58, 146, 207

mp Pendulum Mass. 64–66, 69–71, 87

ms Static Margin. 17, 18, 44

m Aircraft Mass. 40, 41, 43, 45, 48, 49, 58, 118, 137, 153, 176, 183, 191, 207, 208, 225, 229

nx x− wise Load Factor Component in Aerodynamic Reference Frame. 41, 43

ny y− wise Load Factor Component in Aerodynamic Reference Frame. 41

nz z− Load Factor in Aerodynamic Reference Frame. xiv, xv, 21, 41, 43, 46, 47, 111, 114, 117,
119, 122, 127, 132, 151–161, 166–168, 179, 184, 190, 223, 226, 227

nzc Commanded Load Factor. xiv, xv, xvii, 151–163, 165–168, 170, 171, 174–177, 180–183, 185,
187–189, 192, 194, 221, 223–226, 228, 230

nref
z Reference Model for Load Factor. xiv, xv, 151–161, 167, 226

n Number of States. 74

p Roll Rate. 36, 42, 43, 47–49, 58, 59, 164, 168, 179, 184, 190, 227

q Pitch Rate. 21, 29, 36, 42–48, 53, 55, 109, 111, 115–117, 119, 122, 127, 132, 152, 168, 179,
184, 190, 227

rair Perfect Gas Constant. 52

r Yaw Rate. 36, 42, 43, 48, 49, 59, 164, 168, 179, 184, 190, 227

s Laplace Variable. 55, 65, 78

up Pendulum Torque Command. 64, 65, 67, 68, 74, 87

u Command Vector. xv, 75, 76, 87, 110–112, 124, 125, 153, 164, 172, 173, 222

wz Vertical Turbulence in Earth Reference System. 60, 61, 109, 112, 113, 174
Glossary xxix

w Disturbance Vector in H∞ Standard Form. 75–78, 117, 118, 121, 130

y Measured Output Vector. 75

z1 Error on Longitudinal Reference Model. xiv, xv, 151–158, 160, 167, 171, 223, 226

z2 Error on Roll Reference Model. xv, xvii, 163, 165, 167, 168, 171, 174, 175, 181–183, 188,
189, 223, 224, 226, 228

z3 Error on Yaw Reference Model. xv, xvii, 163, 165, 167, 168, 171, 174, 175, 181–183, 188,
189, 223, 224, 226, 228

z Plant Output in H∞ Standard Form. 75–78


Acronyms

CoP Center of Pressure. 17, 96

LoD Lift-over-Drag ratio. 3, 4, 12, 19, 96

AMC Aerodynamic Mean Chord. 44, 51

ASF Acceleration Sensitivity Function. xii–xiv, 77–80, 82, 88–90, 107, 108, 111, 115–118, 120,
121, 123, 124, 126, 128, 129, 133, 136, 150, 151

AVL Athena Vortex-Lattice. 142–146, 217

BLI Boundary Layer Ingestion. 3

BPR By-Pass Ratio. 3

BWB Blended Wing-Body. xiv, xix, 3–7, 11, 12, 14–22, 24, 31, 35, 42, 44, 48–51, 54–56, 58,
64, 83, 87, 93, 94, 96, 98, 99, 102, 104, 106, 108, 115, 117, 130, 139, 140, 142, 143, 145,
147, 150, 177, 186, 193, 199, 201–203, 212

CCV Control-Configured Vehicle. 25–28

CFD Computational Fluid Dynamics. 19, 29, 139, 145, 147, 200

CG Center of Gravity. xi, 4, 17, 18, 22, 25, 40, 42, 44, 52, 56, 57, 87, 96, 97, 102, 178

CO Collaborative Optimization. 28

DOC Direct Operating Cost. 17

EHA Electro-Hydraulic Actuator. 20, 102

EMA Electro-Mechanical Actuator. 20, 102

FbW Fly-by-Wire. 24, 111

FCS Flight Control System. 4, 11, 17, 19–21, 28, 52

FPO Future Projects Office. 50

FSQP Feasible Sequential Quadratic Programming. 29

GW Gross Weight. 26

HQ Handling Qualities. 16

xxxi
xxxii Acronyms

HTP Horizontal Tail Plane. 4, 5, 17, 29–31, 51, 55

HWB Hybrid Wing-Body. 3, 18, 20, 99

ISA International Standard Atmosphere. 50

LFR Linear Fractional Representation. xv, xvi, 7, 139, 145–147, 149, 150, 165, 174, 181, 184,
200, 219, 220

LFT Linear Fractional Transformation. 31

LMI Linear Matrix Inequalities. 5, 30, 31, 78, 84

LQ Linear Quadratic. 23, 29, 74, 75

LQG Linear Quadratic Gaussian. 74

LTI Linear Time-Invariant. 46

MDAO Multidisciplinary Analysis and Optimization. 202

MDO Multidisciplinary Optimization. 5, 18, 24, 28, 29

MEA More Electrical Aircraft. 20

MIMO Multi-Input Multi-Output. 75–77, 134, 162

OAD Overall Aircraft Design. 14, 58

OEI One Engine Inoperative. 17

ONERA Office National d’Etudes et de Recherche Aérospatiale. 7, 19, 145

OWE Operating Weight Empty. 26

PI Proportional-Integral. 21

PID Proportional Integral Derivative. 81

RMS Root Mean-Square. xii, 88, 89, 121, 146, 169

RMT Reference Model Tracking. xv, 149–152, 155, 159, 160, 162, 163, 165, 171

RSS Relaxed Static Stability. 25–27

S&C Stability & Control. 5, 11, 29, 118

SAS Stability Augmentation System. 5, 15, 20, 21, 25, 26

SHA Servo-Hydraulic Actuator. 102

SISO Single-Input Single-Output. 76, 88, 162, 163


Acronyms xxxiii

TLAR Top-Level Aircraft Requirement. 12, 14, 28, 50

TOFL Take-Off Field Length. 50

TsAGI Russian Central Aerohydrodynamic Institute. 14

VMO Velocity Maximum Operating. 52

VTP Vertical Tail Plane. 5, 93, 94, 98, 201

W&CG Weight & Center of Gravity. 56, 94, 96


INTRODUCTION

1
INTRODUCTION 3

Context

When it comes to minimizing aircraft fuel consumption, a direct lever from an aircraft designer
point of view is improving the aerodynamic efficiency of the airframe. For that purpose a first
approach consists in optimizing current aircraft shape, the so-called “tube-and-wing” design, tak-
ing advantage of recent improvements in computational power for aerodynamics and structure
modeling. However it is likely that such an approach is limited to a few percent gains compared
to already flying airplanes, the classical tube-and-wing design having already been improved for
decades. A more ambitious yet demanding strategy consists in departing from traditional air-
craft design in order to conceive a radically new airframe shape, which would potentially achieve
significant fuel consumption gains, and thus meet 21st century economical and environmental
challenges.

Among all so-called “non-conventional” aircraft configurations, few have as much potential as
the flying wing, also known in literature as Blended Wing-Body (BWB) or more recently Hybrid
Wing-Body (HWB) . Indeed an aircraft fundamentally requires four separate functions in order to
safely carry payload airborne: namely Lift, Transport, Control and Propulsion. On a conventional
aircraft each of these four functions are ensured by separate physical components: wing creates
Lift, Transport is ensured by a cylindric fuselage, vertical and horizontal tail as well as control
surfaces on the tails and the wing control the movement of the aircraft with respect to the air, and
Propulsion is ensured by the engines. Such a decoupling has proved very efficient both from an
operational and a conceptual point of view: conception of each physical component is essentially
uncoupled one from another 1 . On a BWB however, Lift, Transport and Control functions are
gathered in a single lifting surface. This essentially eliminates the need for a dedicated fuselage
and empennages, making this aircraft conceptually the simplest one could think of 2 . Potential
benefits are twofold. On the one hand a greater aerodynamic efficiency is expected. This mainly
comes from a significant reduction in wetted area compared to a conventional aircraft, due to the
deletion of the fuselage and empennages. As a result a higher Lift-over-Drag ratio (LoD) is
achieved. On the other hand removing the fuselage and empennages also implies significant mass
savings.

Yet, as often in design, simplicity requires a lot of engineering and ingenuity. This is particularly
true for BWB design, where challenges to overcome are as high as expected benefits. Indeed
what makes this configuration particularly attractive, namely functions gathering and couplings, is
also what makes it so hard to design. For instance, lifting surface design should not be performed
without taking into account passengers accommodation constraints, whereas those two problems
are decoupled in classical aircraft design.

1
Of course in reality coupling do exist, and are closely monitored so that optimum for each component does
not result in a poor global optimum. For instance nowadays’ engines higher By-Pass Ratio (BPR) induce more
aerodynamic interactions between the engines and the wing that may significantly deteriorate the aerodynamic
efficiency. Similarly wing design such as its Cm0 and its x-wise position strongly influence sizing of the empennages.
2
A last step towards integrating functions comes naturally, that is embedding propulsion systems within the
lifting surface. Interestingly this idea came naturally in the early age of jets, see in particular the De Havilland
Comet. Research projects nowadays come back to this intuition when investigating Boundary Layer Ingestion (BLI)
and distributed propulsion technologies, which are however out of scope of our study.
4 INTRODUCTION

Beyond these coupling challenges, aerodynamic-related control specificities of BWB con-


figurations add complexity to the design process. First of all, aircraft control is performed by
non-conventional control surfaces named elevons 3 . These movables spanning the whole trailing
edge act simultaneously on pitch and roll axes. Moreover due to the lack of Horizontal Tail
Plane (HTP) these control surfaces are used both for statically equilibrating the airplane and
dynamic maneuvering. Consequently elevons substitute the horizontal stabilizer – also known as
trim –, elevator and ailerons of a conventional aircraft. As a result a correct sizing of these control
surfaces requires taking into account both static and dynamic criteria, both on longitudinal and
lateral axes.

Secondly, as already mentioned, the BWB lacks a HTP . Trailing edge elevons therefore
feature a smaller longitudinal lever arm with respect to the aircraft Center of Gravity (CG)
compared to the elevator on a conventional aircraft, possibly resulting in degraded pitch authority.
Moreover control surfaces dimensions and shape are tightly constrained by the aircraft glider shape
and profiles, mainly optimized for cruise performance. Thus large control surfaces with significant
hinge dimension are obtained, leading to challenging actuation issues.

Finally among all contradictory requirements at stake during BWB sizing, one is particularly
challenging and yet to solve: how to handle the compromise between aerodynamic performance
and stability. Optimizing the aircraft external shape for aerodynamic performance considerations
– for instance maximizing LoD ratio – indeed tends to induce a longitudinally unstable aircraft,
which may be a threat to security if too strong. More precisely instability is handled by Flight
Control System (FCS) , constantly computing control surfaces deflections in order to keep the
aircraft around its equilibrium position. Any FCS failure or control surfaces reaching their
stop reverts the aircraft to its open-loop unstable behavior, possibly making it uncontrollable and
leading to a catastrophic event. Beyond considering such an adverse eventuality, strong instability
may have a detrimental effect on FCS power consumption and loads, for controls are required
to move permanently and fast to maintain the aircraft around its equilibrium position. Nowadays
recent aircraft tend to be neutrally stable, and sometimes slightly unstable for some specific flight
points. The flying wing considered in our study however is an order of magnitude more unstable.
Yet FCS sizing for an unstable aircraft is still an open problem. More precisely, for checking
that a certain level of instability is acceptable, a stabilizing control law needs to be designed. But
synthetizing such a law in turn requires a fixed aircraft model along with control surfaces and
actuators especially. These two problems are therefore coupled. Whenever controls and actuators
are under-sized, a safety threat exists. If over-sized on the contrary and no power consumption
excess is found, then aerodynamic performance could have been improved.

Today determining an acceptable instability level is still an open question, that slows down de-
velopment of flying wings. Either they are conceived too unstable and at best they are very systems
energy demanding, at worst dangerous, or they are conservatively stable and lack aerodynamic
efficiency.

3
“Elevon” is a contraction of “elevator” and “aileron”. This technology is not new, see for instance Concorde.
INTRODUCTION 5

Previous Studies

In literature, [Saucez, 2013] studied handling qualities of an Airbus long range BWB . An aero-
dynamic model for handling qualities is first developed. Then a list of handling qualities criteria
based on longitudinal and lateral equilibria is established. From this set of criteria it is then
checked that control surfaces efficiency is sufficient in to properly equilibrate and maneuver the
aircraft in the whole flight domain and for different loading scenarios. However instability control
aspects are little addressed, only open-loop handling qualities being considered. Moreover control
surfaces size remains fixed to the initial design assumptions. Finally all criteria are equilibria, not
dynamic maneuvers.

In a different domain, [Perez et al., 2006] incorporated a Stability & Control (S&C) module
into a Multidisciplinary Optimization (MDO) process. It is shown that for a conventional aircraft
configuration there are benefits to expect when relaxing open-loop stability. More precisely two
optimizations are run in order to geometrically size a tube-and-wing aircraft, respectively without
and with incorporating a Stability Augmentation System (SAS) to fulfill handling qualities
constraints. In the optimization process, control law gains – for the augmented aircraft – as
well as geometrical parameters, such as wing planform and position, Vertical Tail Plane (VTP)
and HTP sizes as well as control surfaces area are computed. Handling qualities constraints are
expressed both in terms of maximum authority to trim the aircraft and maneuvering in the whole
flight envelope, as well as constraints on open- and closed-loop poles characteristics. Results show
an improved performance of the augmented aircraft in terms of maximum achievable range, mainly
due to a gain in wetted area enabled by HTP and VTP downsizing. However the augmented
aircraft static margin remains positive – ie the aircraft is stable. Moreover this process is applied
to a conventional aircraft and not a BWB , and the gain structure is a state-feedback law.

Eventually in Control community a research field named “Plant-Controller Optimization”,


“integrated design and control”, or “codesign”, consists in simultaneously computing a controller
and some physical parameters. More precisely in aeronautics [Niewoehner and Kaminer, 1996]
simultaneously optimize a longitudinal controller and the HTP area. Their approach uses the
H∞ theory solved with the Linear Matrix Inequalities (LMI) framework. The major interest
of this study is that it translates handling qualities constraints for HTP sizing into frequential
constraints cast as H∞ norms. However, besides being applied on a conventional aircraft for
longitudinal sizing only, a drawback of such approach is that the resulting controller is full-order,
that is possibly embedding high-order internal dynamics. Yet controllers in the aeronautics industry
usually feature a fixed structure, coming from engineering knowledge on systems to be controlled.
Keeping such a predefined structure within a control optimization process is a feature of much
interest for the industry.

Summary on Existing Work

In a nutshell, previously mentioned studies demonstrate that there are ways to simultaneously
optimize a controller and physical parameters, either using “MDO” or “Control” approach. Nev-
ertheless:
6 INTRODUCTION

• Industrial constraints on controller structure are not fulfilled.

• Most studies focus only on longitudinal sizing.

• Control surfaces consist mainly in a conventional elevator.

• More generally, integrated design and control has not been addressed for BWB sizing.

This is what justifies our study. It consists in concurrently optimizing control surfaces,
actuators and flight control laws taking into account longitudinal and lateral instabilities
as well as industrial structure for controllers, for unstable configurations such as the
BWB .

Scientific Process

Our study consists first in elaborating a handling qualities procedure for sizing an unstable aircraft.
More precisely our objective is to formalize the interactions between different scientific disciplines,
such as aerodynamics, open and closed-loop handling qualities, and systems, and their associated
constraints. For that purpose all sizing disciplines are first identified, together with the sizing
parameters they influence and the constraints they require. These disciplines are then sorted on
a time scale, in order to provide guidelines to aircraft designers when sizing an unstable aircraft.
This first process is called sequential, or iterative. Then two alternative processes are proposed,
which are shown to reduce development time. The first one consists in a combined optimization
of control surfaces and flight control laws, instead of a sequential design. The second one is a
combined optimization of actuators bandwidth and flight control laws. These different processes
are extensively explained in chapter 4.

The next part of our study is devoted to numerically setting up the first alternative process
proposed in chapter 4: that is sizing control surfaces actuators bandwidth with constraints on
longitudinal stabilization. For that purpose a combined optimization of a longitudinal stabilizing
control law and actuators bandwidth through structured H2 /H∞ synthesis is performed. More
precisely the objective consists in simultaneously optimizing actuators bandwidth of trailing edge
elevons together with a longitudinal stabilizing control law, using a weighting scheme on accelera-
tion sensitivity function. To achieve this, a synthesis scheme weighting the acceleration sensitivity
function is developed, to synthesize a stabilizing control law guaranteeing adequate handling qual-
ities. A main interest of this scheme is that it uses the H∞ framework, that allows working with
structured synthesis algorithms capable of simultaneously optimizing actuators bandwidth and
control gains. Then parameterized bandwidth are added to the structured synthesis, to enable
their optimization together with a stabilizing control law. Finally an energy minimization criterion
for controlling the airplane is written. The aim of this work is to find the optimal bandwidth of
actuators guaranteeing a minimal energy consumption while stabilizing the aircraft. This study
is explained in details in chapter 5 and led to two publications [Denieul et al., 2015a] [Denieul
et al., 2015b].
INTRODUCTION 7

A next phase is devoted to setting up a longitudinal-lateral aerodynamic model parameterized


by control surfaces size. The objective is to obtain a state-space representation parameterized by a
continuous parameter representing total control surfaces span on the outer wing trailing edge. For
that purpose a parameterized geometrical model of the aircraft is developed, knowing that later
the final goal is to optimize this parameter representing control surfaces size. Then control surfaces
aerodynamic efficiencies are computed for several geometrical configurations, both in longitudi-
nal and in lateral-directional, because trailing edge elevons act on all those axes. Aerodynamic
computations are run with linear low-fidelity method and then calibrated using higher-fidelity data
available on baseline BWB . Finally discrete state-space representations obtained are approxi-
mated with polynomials in order to get a continuously parameterized state-space representation
suitable for further continuous optimization. To do so methods developed at Office National
d’Etudes et de Recherche Aérospatiale (ONERA) were used that compute polynomial approxi-
mations into a Linear Fractional Representation (LFR) representation, a well adapted formalism
for structured synthesis optimization algorithms used hereafter. This procedure is developed in
chapter 6.

The final step consists in simultaneously optimizing multi-axes control laws and control surface
sizes under stabilization and maneuvers constraints.The objective is to get in a single optimization
loop both control law gains on all axes and optimal control surfaces sizes that fulfill longitudinal
and lateral handling qualities constraints. For that purpose a control synthesis scheme that allows
for a simultaneous computation of fixed structure control law gains on all axes is developed. Then
constraints that dynamically size control surfaces are expressed as constraints on H∞ norms of
transfer function in order to be easily tractable by the optimizer. Finally the optimization process
is run on the whole flight domain in order to get sizing point for control surfaces. This final study
and results are explained in detail in chapter 7.
Part I

STATE-OF-THE-ART

9
Chapter 1

State-of-the-Art

“What has been will be again, what has been done will be done again. There is nothing new
under the sun.”
- Ecclesiastes 1:9

Contents
1.1 Design Challenges of Commercial Blended Wing-Body Aircraft . . . . 11
1.1.1 Overall Aircraft Design Aspects . . . . . . . . . . . . . . . . . . . . 12
1.1.2 Stability and Control Challenges . . . . . . . . . . . . . . . . . . . . 14
1.1.3 Flight Control System Sizing for Blended Wing-Body . . . . . . . . . 19
1.2 Interactions of Aircraft Design and Control . . . . . . . . . . . . . . . 24
1.2.1 Research in Control-Configured Vehicle . . . . . . . . . . . . . . . . 24
1.2.2 Multidisciplinary Optimization Approach . . . . . . . . . . . . . . . . 27
1.2.3 Control-Based Approach . . . . . . . . . . . . . . . . . . . . . . . . 29
1.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

This part is devoted to analyzing prior art to our study. Whereas much relevant literature
has been published concurrently with our work, showing a more than ever active research field
of studying interactions of design and control especially for BWB , work published since 2014
is is voluntarily not included in this part. The rationale of this state of the art review is indeed
to justify choices and orientations of our study based on existing work at the start of the study.
However relevant studies that fed our reflexion are included in the Scientific Contribution part,
which the reader is referred to for more information on present research publications.

This literature review is divided into two separate topics: first in section 1.1 design challenges
for BWB are explained, with a particular focus on S&C issues and impacts on FCS sizing. Then
in section 1.2 the general problem of simultaneous design and control is developed. Aeronautical
cases are studied, and two concurrent methods are examined.

1.1 Design Challenges of Commercial Blended Wing-Body Aircraft

While not recent – see for instance the Horten H III f glider built in 1938 on figure 1.1(a) and the
Northrop N-1M from 1940 figure 1.1(b) –, the flying wing configuration or its commercial aircraft
counterpart BWB 1 represents a potential revolutionary step towards more efficient aircraft in
1
Strictly speaking, flying wings differ from BWB in that the former features straight leading and trailing edge,
with no structural difference between payload transportation and lifting surface, while the latter is composed of a

11
12 Chapter 1. State-of-the-Art

terms of fuel consumption, operation costs and noise reduction. A brief overview of where these
gains come from is given in section 1.1.1, and a deeper focus is set on yet to be treated stability
and control challenges associated to this unconventional configuration.

(a) Horten H III f glider (Source: Smithsonian NASA (b) Northrop N-1M (Source: Smithsonian NASA
Museum.) Museum.)

Figure 1.1: Flying wings prototypes.

1.1.1 Overall Aircraft Design Aspects

1.1.1.1 Aerodynamic Performance

The BWB concept was theoretically developed by Robert Liebeck in [Liebeck et al., 1998]. At
that time he was an aerodynamics engineer at the MacDonnell-Douglas company, which then
became part of the Boeing company. The concept emerged from a question by NASA about
the most promising aircraft configurations for subsonic commercial airplane (see also [Chambers,
2005]). To answer this demand a conceptual study on how to accommodate passengers in a flying
body while minimizing the wetted area Swet of this body emerged. For a fixed volume required
for passengers accommodation, minimum wetted area is a sphere. Streamlining this sphere leads
to two concurrent shapes, which after integration of the wing, propulsion and controls become
the conventional tube-and-wing shape and BWB respectively. Better integration of transport,
lift, controls and propulsion for the BWB configuration leads to a decrement of 33% wetted area
compared to tube-and-wing design for the same Top-Level Aircraft Requirements (TLARs) of 800
passengers carried on a 7000NM mission. As maximum cruise LoD is linked to the wetted area
aspect ratio λ = b2 /Swet , a significant improvement of 15% maximum LoD is achieved. This
remarkable figure has been confirmed in other flying wing preliminary studies [Martínez-Val and
Pérez, 2005; Liebeck, 2004; Qin et al., 2004; Bolsunovsky et al., 2001]. Conceptual process that
led to the BWB concept is illustrated on figure 1.2 and the Boeing BWB is visible on figure 1.3.

Performance gains of BWB configuration mostly come from a reduced wetted area
Swet .

flattened centerbody and an outer wing, with usually a kink between both. See [Martínez-Val et al., 2010] for more
details.
1.1. Design Challenges of Commercial Blended Wing-Body Aircraft 13

Figure 1.2: Conceptual view on wetted area gains of a BWB vs conventional aircraft. Source:
[Liebeck, 2004]

1.1.1.2 Structure

A consequent challenge arising while defining and evaluating this configuration is that of structure
sizing. In conventional tube-and-wing design, the cylindric shape of the fuselage is well tailored
to handle radial pressure efforts stemming from cabin pressurization. How to efficiently conceive
a flattened fuselage resisting pressurization loads is still an open matter. In [Liebeck et al., 1998]
it is this constraint removal which then directs the design process towards blended wing-body
solution, while “assuming that an efficient structural concept could be developed” remains to
be treated. A debate exists between relative efficiencies of two structural concepts: simple skin
sandwich shell versus double skin vaulted solution, see figure 1.4. In the simple skin solution,
all pressurization loads, but also bending and torque, are withstood by the outer aerodynamic
planar panel. According to [Martínez-Val and Pérez, 2005], resisting internal pressure with planar
panels being very inefficient, this solution would lead to undesirable extra weight. However this
concept was selected in [Liebeck, 2004]. In the opposite, double-skin vaulted solution divides
structure among inner cylindrical structure, rather similar to conventional fuselage, withstanding
pressurization loads. An outer skin bears conventional loads as a wing. A potential threat to this
concept would be the case of a rupture of the inner skin, that would require the external panels
to be sized for pressurization loads also, removing the benefits of the concept [Martínez-Val and
Pérez, 2005]. Finally [Bradley, 2004] suggests abandoning the double-skin concept in favor of an
adequately strengthened outer skin withstanding all pressurization as well as bending loads.
14 Chapter 1. State-of-the-Art

Figure 1.3: Boeing BWB concept. Source: [Liebeck, 2004]

1.1.1.3 Capacity and Operations

Several TLARs were investigated in past BWB studies, ranging historically from very large aircraft
to more conventional capacities. In the 1990s Airbus studied together with the Russian Central
Aerohydrodynamic Institute (TsAGI) a Very Large Aircraft (VELA) featuring 750 passengers in
3-class layout, and 7400NM range [Bolsunovsky et al., 2001]. In the same order of magnitude
first Boeing studies in [Liebeck et al., 1998] featured 800 passengers spread on a double-deck,
and a 7000NM range. Configurations then converged towards more conventional requirements:
[Liebeck, 2004] studies a 450 passengers, 7500NM range aircraft. Finally in [Martínez-Val et al.,
2010], potential impacts of BWB in air traffic and airport operations is studied. Significant
benefits are found in terms of airport congestion, community noise, air traffic altitude spreading,
and emissions.

1.1.2 Stability and Control Challenges

After a brief overview of the design potential and issues of the BWB configuration from an
Overall Aircraft Design (OAD) perspective in previous part, this section emphasizes control-related
challenges implied by BWB design. Open-loop handling qualities are treated in section 1.1.2.1,
then trade-off between stability and cruise performance is examined in section 1.1.2.2.
1.1. Design Challenges of Commercial Blended Wing-Body Aircraft 15

(a) Double-skin vaulted shell. (b) Flat sandwich shell.

Figure 1.4: Structural concepts for BWB pressurized shells. Source: [Liebeck, 2004].

1.1.2.1 Handling Qualities

In [Bolsunovsky et al., 2001], stability and control challenges are quickly raised. Low longitudinal
and lateral stability, as well as small natural yaw damping, are shown to have been the weak point
of flying wings since the beginning of aviation. Control is ensured through full trailing-edge span
elevons and split rudders. Multi-functional surfaces are then quoted as highly beneficial to control:
antisymmetrically deflecting the split rudder acts as drag device, and elevons are used both for
longitudinal and lateral control, as well as flaps and trim devices. Aerodynamic effectiveness of
control surfaces is evaluated in subsonic wind-tunnel tests. It is found that stall recovery imposes
high requirements for control surfaces actuation rates: ≈ 50˚/s for elevons and ≈ 100˚/s for split
rudders. However no comprehensive study is performed on control surfaces sizing, neither for
equilibrium nor for dynamic maneuvers.

[De Castro, 2002] specifically studied flying and handling qualities of a fly-by-wire BWB
civil transport aircraft. Both longitudinal and lateral handling qualities are studied on several
BWB configurations, and piloted evaluations on a flight simulator are performed. Concerning
longitudinal behavior, target static margin, i.e. level of longitudinal instability, is found to be
critical in the design of the airplane: an optimum should be found due to low control authority.
More precisely, a too unstable aircraft results in a huge requirement for control power, even
possibly questioning flight safety. On the contrary, too much positive static margin entails a poor
maneuverability. These phenomena, while well known on conventional aircraft, are particularly
acute for BWB due to its low control authority. The prescribed solution is a slightly unstable
airplane; however such advice is only qualitative and no trade is performed. As a consequence a
SAS is required. In this study a pitch rate command / attitude hold system is used for longitudinal
motion control. A large part of the study also focuses on directional stability issues, that are found
to be notably problematic for the studied configurations. This comes from the lack of vertical
surfaces, implying an unstable dutch roll mode that requires high gains to be controlled.

Then in this study no control allocation for a multi-control use of control surfaces is used: only
16 Chapter 1. State-of-the-Art

conventional effectors, such as elevator, ailerons and rudder, are considered. This may explain
conclusions on the lack of control authority. Also while evaluating different BWB configurations,
no trade is performed on control surfaces or control power sizing, while acknowledging a need for
an optimization.

[Chudoba, 2001] developed a generic handling qualities tool to assess control authority in
conceptual design for unconventional aircraft such as flying wings and non-symmetric airplanes.
A control allocation module is developed but is limited to optimal trim computation in cruise from
drag minimization point of view. The structure of feedback control law is limited to angular rates
feedback; control law gains are computed incrementally so as to increase stability until a prescribed
stability requirement is reached. This approach is suited to stable aircraft, even slightly. The
purpose of this study is more to estimate control power, rather than assessing handling qualities
issues. This approach is also developed by the authors in [Coleman and Chudoba, 2007].

Figure 1.5: Airbus BWB concept. Source: [Saucez, 2013]

In [Saucez, 2013] a comprehensive assessment of Handling Qualities (HQ) issues is performed.


This study is a PhD thesis conducted together with Airbus Future Projects Office, aiming at
assessing feasibility of a 450 passengers long-range BWB . First handling qualities issues of the
BWB configuration are summarized:

• Longitudinal and directional instability

• Take-off rotation issues

• Lack of roll control authority

• Lack of directional control authority

To solve these challenges, a process is developed that takes advantage of BWB geometric
specificities, namely its large lifting surfaces and chords, that provide degrees of freedom for
aerodynamic design. Take-off and landing performance are also investigated. It is shown that
no major show-stopper threatens the feasibility of this configuration, provided vertical surfaces
are sized to provide a sufficient open-loop dutch roll damping. Another conclusion is the need
for multi-control surfaces, namely elevons acting both as roll and pitch devices, largely improving
1.1. Design Challenges of Commercial Blended Wing-Body Aircraft 17

control authority compared to the “mono-control” baseline. Also a strong pitch instability is
demonstrated, that requires robust stabilizing control laws. In this study only pitch damper is
investigated, showing an adequate improvement of stability issues. However only equilibria and
trim conditions are investigated, no dynamic simulation being taken into account for control
surfaces sizing. Also no optimization is performed on FCS appart from vertical surfaces sizing;
the approach focuses on evaluating different architectures for yaw control and their impact on the
mission. Then only simple backup rates feedback control laws are included. Our work is following
this study, taking this PhD thesis conclusion as our starting point and optimizing the proposed
BWB configuration taking into account control laws.

Static equilibria for engine failure case have been performed in a separate study in [Saucez and
Boiffier, 2012]. Three unconventional yaw devices are evaluated for compensating yaw moment
induced by One Engine Inoperative (OEI) operations: crocodile ailerons, thrust vectoring, and
winglets with rudders. Relative efficiencies of proposed solutions are evaluated through a Direct
Operating Cost (DOC) criterion. The most promising solution is a combination of winglets and
non-permanent thrust vectoring. This study focuses on lateral balancing only, no dynamic control
law being included.

1.1.2.2 An Aerodynamic Specificity: Trade-off between Stability and Cruise Perfor-


mance

A major challenge when designing a flying wing concerns the trade-off between cruise performance,
trim and pitch stability. In other words how to conceive an aerodynamic shape which, for a given
CG range, is equilibrated in cruise, thus requires no trim deflection leading to trim drag, and also
statically stable2 meaning that perturbations from the equilibrium will not lead to strong pitch
departure, without compromising aerodynamic efficiency? Such a trade-off is made difficult due
to the absence of HTP .

Notionally these conflicting requirements involve two remarkable aerodynamic points: the
Center of Pressure (CoP ) Xcp and the aerodynamic neutral point XF . For a single lifting
surface:

• Xcp is the point where moments created by aerodynamic forces equal zero. When the
Center of Gravity XCG is located at this very point, the aircraft is equilibrated.

• XF is the point where moments created by aerodynamic forces are constant with respect
to angle-of-attack α . Linear aerodynamic theory for lifting surfaces shows the existence
of such a point, for which the aerodynamic pitching moment equals Cm0 whatever α .
If XCG lies forward XF , then the aircraft is said statically stable. Static margin ms is
defined as ms = XF −X l
CG
. A positive static margin is therefore equivalent to a statically
longitudinally stable aircraft.

For a comprehensive description of these points, see [Taquin, 2009] and [Boiffier, 1998]. It can
2
Or at least not too unstable.
18 Chapter 1. State-of-the-Art

be shown that Xcp and XF are linked through the relation:

Cm0
Xcp = XF − l (1.1)
Cz
Xcp and XF are oriented positive towards the rear of the aircraft. Cz is a function of α
, therefore Xcp moves with α whereas XF does not3 . In classical airfoils design Cm0 is
usually negative.

As already emphasized, longitudinally balancing the aircraft requires XCG = Xcp . Then with
the assumption of Cm0 < 0, equation (1.1) clearly shows that case results in ms < 0. This
proves at first sight that a single lifting surface with a negative zero-lift pitching moment cannot
be both balanced and stable, hence the need for an empennage, which flying wings do not have.
Balancing and stabilizing a flying wing therefore requires a careful design on chord-wise but also
span-wise lift distribution, playing on 3D and sweep effects. How this fundamental topic is handled
in literature is discussed hereafter.

According to [Raymer, 1999], "The Hortens (see figure 1.1(a)) employed a design philosophy
of reducing the lift at the wing tips to nearly zero, twisting the wing to generate most of the lift
on the inboard part of the wing. This allowed moving the CG forward and created a design very
stable in pitch [...]. The inefficiency of the unusual lift distribution was corrected with increased
aspect ratio. Notionally, you can think of the Horten wings as conventional aft-tailed aircraft but
with those aft tails stuck out on the wing tips."

In the same book, a different approach is also proposed: “Alternatively, a "reflexed" airfoil
can be selected, having the trailing edge lifted slightly to provide a naturally stable airfoil. Such
airfoils tend to be less efficient and are typically limited to slower aircraft.” Here stable airfoil
means Cm0 > 0. We see that both solutions imply penalty on aerodynamic efficiency, that have
to be traded against penalty on trim drag for performance-optimized profiles.

The second solution seems to be favored in [Liebeck, 2005]. On the initial BWB design, aft-
cambered supercritical airfoils are assumed for the outer wing for transsonic performance, whereas
centerbody uses a reflexed airfoil for pitch trim. However in the same study, several conflicting
requirements are detailed that challenge the airfoil design. First, cruise deck angle requires that lift
is generated at a low angle of attack, implying aft cambered profile on centerbody airfoil, with a
strong nose down zero lift pitching moment. On the contrary as already discussed trim and static
stability impose to minimize this nose-down pitching moment and the use of aft-cambered profiles.
In return, “if the BWB is designed with negative static margin (unstable), it will require active
flight control with a high bandwidth, and the control system power required may be prohibitive.”

More recently with the advent of MDO capabilities, aerodynamic glider optimization for
BWB shape got investigated a lot. Some of these studies involved constraints on trim and static
stability. A main contribution in this field is provided by [Sargeant et al., 2010], in which the
design of a trimmed, statically stable and efficient HWB is discussed. First the authors recall the
two previously mentioned techniques – namely using twist together with washout twist in order to

3
Still in linear aerodyamics theory.
1.1. Design Challenges of Commercial Blended Wing-Body Aircraft 19

download the wing tips, or downloading the aft part of centerbody using reflex-cambered stable
profiles – and their associated drawbacks. They then claim to achieve a trimmable and statically
stable configuration while using supercritical efficient profiles on the outer wing, thanks to front-
loaded profiles on the centerbody, the so-called “leading-edge carving”. The advantage would be
not to download too much the centerbody aft part, and to “normally” design the outer wing. A
main contribution of this study is the evidence through Computational Fluid Dynamics (CFD)
experiments of the strong influence of 3D aerodynamic effects on the stability analysis. Indeed
a 2D analysis on their configuration would lead to the false conclusion of an unstable aircraft.
Another claimed advantage of leading-edge carving is a near-elliptical lift distribution for a wide
range of angle of attack, still thanks to 3D effects.

In a similar topic, [Mader and Martins, 2013] studied the aerodynamic optimization of flying
wings with constraints on stability. Several computations are carried out: lift constrained drag
minimization on a rectangular wing, addition of bending moment, static and dynamic stability
constraints. Variables are both the planform and profiles. Several previously mentioned phenomena
are found again through optimization: Cm0 may be modified through profile camber changing,
or by twisting the wing tips if the wing has sweep. The aerodynamic center can be moved by
modifying sweep. A conclusion is that balancing and stabilizing the aircraft can be achieved
through profile camber or wingtip twisting with sweep. As sweep is necessary for high speed, it is
natural to use it. This last assumption is quite contradictory to previous studies. It is also found
that adding stability constraints in the optimization process alters the achieved optimum in terms
of drag counts.

A last study of interest, [Meheut et al., 2012], was performed by ONERA on the same
baseline as [Saucez, 2013]. An aerodynamic optimization is performed for LoD maximization,
with planform and profiles variables. The main interest of this work is to include a constraint on
a maximal negative Cm0 coming from take-off rotation issues. A penalty on LoD results of
including this constraint. Results also indicate that contrary to what linear aerodynamic theory
says, for a given planform there is an influence of profiles on XF location. This should probably
come from high-speed non-linearities.

As a conclusion for this part, design of a stable and cruise efficient flying wing is still an
open issue, but significant improvements were made recently thanks to the advent of aerodynamic
optimization tools. A question that also arises is: isn’t a constraint on static stability too con-
servative? Perhaps a slight instability should be allowed, relaxing constraints on the aerodynamic
design, with not too much penalty on FCS design. FCS sizing is therefore investigated next.

1.1.3 Flight Control System Sizing for Blended Wing-Body

FCS sizing is a key challenge for BWB , mainly due to unconventional flight dynamics associated
to multiple redundant control effectors. This section is devoted to reviewing state-of-the-art on
this topic.
20 Chapter 1. State-of-the-Art

1.1.3.1 Control Surfaces and Actuators Sizing Problem

In [Roman et al., 2000], design challenges associated to secondary power for control surfaces
actuation are discussed. These control surfaces have a small lever arm with respect to longitudinal
center of gravity, hence the need to incorporate multiple redundant, fast moving control surfaces.
They are also multi-functional, including trim, longitudinal and lateral control, pitch stability
augmentation, and wing load alleviation. Also the significant so-called “square-cube law” for
hinge moments estimations is introduced. Hinge moments are related to control surface area as
follows: when geometrically scaling an aircraft with a factor λ, the control surface areas increase as
the square of the scale λ2 , whereas hinge moments increase with the cube of this scale λ3 . Large
BWB control surfaces lead to high hinge moments requirements. Maximum hinge moments in turn
size the hydraulic system. Furthermore as already mentioned an unstable BWB would require a
SAS causing high deflection rates. And “once the hydraulic system is sized to meet the maximum
hinge moments, the power required is related only to the rate at which the surfaces move.” As
a consequence, combination of high hinge moments and significant deflection rates imply that “
secondary power required can easily exceed that currently available from turbofan engines” and
therefore ‘may become prohibitive.‘” This work concludes by stating that this problem of properly
sizing FCS for a BWB remains to be treated.

[Liebeck, 2005] develops similar considerations. The question of control system bandwidth is
raised: designing a BWB “ with a negative static margin would require active flight control with
a high bandwidth, and the control system power may be prohibitive”.

A comprehensive view on applications of control theory to airplanes is given in [Abzug and


Larrabee, 2005]. A whole chapter is devoted to SAS for naturally unstable aircraft, which is called
by the authors “superaugmentation”. A particular emphasis is set on the necessity for high rates
actuators. Moreover unstable configurations should be sized so that disturbance levels should not
make the actuators saturate (in rate or position), otherwise the aircraft turns unstable: "Control
system rate or position saturation can be particularly deadly for superaugmented airplanes. Once
the control surfaces are operating at actuator-limited rates or against the surface stops, the design
reverts back toward the unaugmented (unstable) case. In design studies, one must identify the
command or disturbance levels that could cause unacceptable divergences due to control rate
and position saturation." A drawback of high bandwidth actuators is also mentioned, that could
limit the design of unstable aircraft: "a limiting factor (for relaxed static stability) is that the
higher bandwidth actuators needed to cope with unstable airplanes tend to interact with the lower
frequency flexible modes."

Few works on the field of actuators and FCS sizing for unconventional configurations have
been published for years, until a renewed interest highlighted by the work of [Chakraborty et al.,
2013b] [Chakraborty et al., 2013a]. This work aims at designing and optimizing control surface
actuators for More Electrical Aircraft (MEA) , first applying the proposed methodology on a
Boeing 737-800 but claiming further research to apply it on a HWB . Actuators models for mass,
volume, kinematics and power consumption are developed; Electro-Mechanical Actuators (EMAs)
and Electro-Hydraulic Actuators (EHAs) are considered. Hinge moments are computed according
to a methodology proposed in [Roskam, 1985c], and an optimization is run in order to minimize
1.1. Design Challenges of Commercial Blended Wing-Body Aircraft 21

FCS mass while sustaining actuation loads on the whole flight enveloppe. However criteria are
only static, no dynamics or control law is considered.

While being pointed out for years as a major challenge for BWB design, actuators and
control surfaces sizing problems for unstable BWB remain largely absent in literature.

1.1.3.2 Review on Control Laws used for Blended Wing-Body Control

In this section a brief overview of longitudinal control laws techniques for BWB is given. Not
much literature exists on the topic, obviously because no commercial BWB or flying wing has
ever flown. Military flying wings, such as the B-2 bomber or Dassault NeurOn do fly, however
information on these programs is classified.

According to [Abzug and Larrabee, 2005], two concurrent solutions exist for pitch augmenta-
tion: the first one consists in a feedback of angle of attack α , pitch rate q and airspeed Vair
. However measures of α and Vair in case of turbulence are likely to be noisy. Therefore the
authors suggest considering inertially based signals instead of air data, that is pitch rate q , pitch
attitude θ and vertical load factor nz .

In [De Castro, 2002] SAS is composed of state-feedback of α , q , θ and Vair .


Eigenvalues assignment techniques compute gains in order to obtain desired closed-loop behavior.
For demanding piloting tasks such as approach and flare, an alternative control law is designed
on reduced two-states [α, q] model. Angle of attack is fed back to provide artificial pitch stability,
and a Proportional-Integral (PI) compensator acts on pitch rate error – pilot commands pitch
rate –. These structures of control laws are found to provide adequate flying characteristics, both
theoretically and through pilots evaluations on simulator. Only flare maneuver reveals delicate
because of the unusual rate-command response type from pilot perspective.

Much significant work in flight control laws for tailless aircraft and BWB is about subscale
demonstrators. [Brinker and Wise, 2001] [Wise et al., 1999] studied reconfigurable flight control
laws for X-36 demonstrator, visible on figure 1.6(b). Dynamic inversion control laws with explicit
model following are used. Robustness issues are pointed out in [Brinker and Wise, 1996]. Then in
[Brinker and Wise, 2001] [Wise et al., 1999] on-line neural networks are implemented to adaptively
regulate tracking error between model reference and actual vehicle dynamics. This neural network
is then capable of faulty actuators detection, and reallocating commands to remaining actuators.
Flight tests of this reconfigurable control law are then carried out, showing improved handling
qualities under actuator failures. Similar research was conducted on the X-48B BWB sub-scale
demonstrator, visible on figure 1.6(a). [Jackson and Buttrill, 2006] presents flight control laws
used for a free-flight model in wind-tunnel tests. Angle of attack α , pitch rate q and average
nacelle effector pressure are fed back to the compensator, for stabilization, damping, and thrust
pitch compensation, respectively. Gains are computed iteratively, using root-locus techniques and
Nichols diagrams for robustness assessment. l1 −adaptive control was also investigated in [Leman,
2009] – see [Hovakimyan and Cao, 2010] for a comprehensive theory on l1−adaptive control –, and
[Goldthorpe et al., 2010] suggests that dynamic inversion was flight-tested, especially in non-linear
22 Chapter 1. State-of-the-Art

flight domain regions such as high angle of attack and sideslip flight points.

(a) X-48B low-speed sub-scaled BWB demonstra- (b) X-36 tailless aircraft demonstrator.
tor. Source: [Liebeck, 2015]

Figure 1.6: Unmanned tailless aircraft.

1.1.3.3 Control Allocation Strategies

Control allocation is the problem of converting equivalent orders into control orders when there
are more effectors than axes to control. A comprehensive survey of control allocation methods
can be found in [Johansen and Fossen, 2013]. This problem is of particular importance for BWB
control, for control surfaces are at the same time redundant – several control surfaces perform
the same function – and multi-functional – namely elevons act as pitch, trim and roll effectors–.
Moments control implies three equivalent orders in pitch, roll, and yaw respectively, that need to
be allocated to all control surfaces. Taking into account position saturations, there may exist an
infinity, a single, or no solution depending on the magnitude of the desired equivalent order. The
problem of finding the “best” order for each control is that of control allocation.

[Cameron and Princen, 2000] comprehensively details control allocation requirements for BWB
.The studied configuration features twenty control surfaces. Of particular interest is the discussion
concerning multi-axes sizing of control surfaces: “The BWB preliminary design process must
simultaneously recognize the control power requirements in multiple axes, a complexity that is
not usually considered in conventional large transport aircraft.” Design of the allocator becomes
critical from an aircraft design point of view, for it will determine pitch, roll and yaw capability of
the aircraft, and therefore imply constraints for instance on longitudinal CG limits, or location
of the main landing gear. Then several requirements are developed concerning the choice of
the allocation algorithm. The authors emphasize the necessary simplicity of the allocator, for
development as well as certification issues. It is also said that “control allocator shall not contain
optimization algorithms”, because no proof of convergence within computational sampling rate
can be made. Therefore the allocator shall be deterministic. This remark is specially important,
for as it will be seen next, several studied allocation algorithms feature optimization procedures.
Finally several good practices are detailed, such as using more effective control surfaces first, and
moving all control surfaces together instead of one by one because it implies less deflection and
rates. However this article only provides guidelines, but no proposal on specific algorithms is made.
1.1. Design Challenges of Commercial Blended Wing-Body Aircraft 23

The simplest method, used in [Saucez, 2013], is called ganging control allocation. It consists
in allocating different priorities for all control surfaces, for instance according to their relative
efficiency on each axis, and to allocate commanded orders on higher priority first. When some
controls are saturated, next priority controls are moved and so on. This method is purely deter-
ministic. Its drawbacks are a lack of flexibility, and more importantly it implies high deflections
and deflection rates. It can also be shown that the entire capabilities of the control architecture
are not achieved: some orders would be physically achievable but are not found by the allocator.

An alternative control allocation strategy, named direct control allocation, was proposed by
[Durham, 1993], and then developed and refined by the same author in [Durham and Bordignon,
1995] [Durham, 1997] [Durham, 2001]. This geometrical method works for linear effectors with
minimum and maximum boundaries, and is based on the concept of attainable moments subset:
it is the space of moments physically achievable by the control configuration. It is proved that
this space is a closed convex zonoedron, and that each point on the boundary of this set is
attained by only one control configuration. Based on this observation a unique allocation is
found whatever the desired moments. Due to its geometrical nature, this technique does not
involve any optimization routine; moreover it guarantees using the whole capability of the control
architectures. Finally minimal deflections are ensured. However computation of the attainable
moments subset is delicate, and could not yet be embedded into a flight computer.

Other methods involve optimization-based control allocation. Over the set of solutions
that satisfy constraints on deflections, solution that minimizes an objective criterion is chosen.
Such objective may be drag [Buffington, 1997], loads [Bodson and Frost, 2011] or maximum
deflection [Burken, 2001] minimization. If the constraints are linear, linear programing techniques
may be applied [Bodson, 2002], speeding up the resolution. Moreover for such cases proofs that
any feasible solution will be found do exist. Equivalence of l2 −optimization approach with Linear
Quadratic (LQ) optimal control is shown under some hypotheses in [Harkegard and Glad, 2005]
[Harkegard, 2004] . However as already discussed embedding optimization algorithms into flight
computers is not yet considered, at least on commercial. For this reason optimization algorithms
such as FSQP [Lawrence et al., 2010][Johansen et al., 2004] were tried on our study and then
discarded.

A last approach derived from optimal control allocation, named pseudo-inverse control allo-
cation is described in [Bodson, 2002] and references therein. Indeed solutions of the l2 −optimal
error minimization problem can be directly computed with the Moore-Penrose pseudo-inverse.
However saturations are not taken into account in this solution. A derived approach, named
redistributed pseudo-inverse control allocation consists in limiting optimal deflections com-
puted using pseudo-inverse to their saturations if they are exceeded, and then compute again the
pseudo-inverse solution for remaining controls. This loop is repeated as many times as there are
controls available. This solution has several advantages. First it does not involve any optimiza-
tion procedure, and the number of loops is deterministic. Then obtained solutions are “clever”,
in the sense that more effective controls are used first. Finally the linear version of this algorithm
makes it easy to include into a control law synthesis scheme. Drawbacks often cited in literature
involve physically attainable moments that cannot be captured using this algorithm, but were not
observed in our study.
24 Chapter 1. State-of-the-Art

For those reasons linear pseudo-inverse control allocation is used in our study for control
law synthesis, and redistributed pseudo-inverse is used for simulation.

To conclude this part, let us see a few concrete implementations of control allocation. [Bor-
dignon and Bessolo, 2002] describes how redistributed pseudo-inverse algorithms are implemented
on the X-35B experimental aircraft. The author’s conclusion are similar to ours, that is “this
method is not truly optimal and thus cannot capture 100% of the control power, but is very
efficient given throughput and storage constraints.” [Goldthorpe et al., 2010] describes control al-
location on the X-48B, featuring some kind of ganging method with tabulated tables to avoid any
online computations. Finally [Waters et al., 2013] implements different control allocation strate-
gies, namely l2−optimal, linear programing, pseudo-inverse and direct allocation, on a BWB .
Wind-tunnel tests are conducted in order to measure difference between desired moments and
actually achieved aerodynamic moments using each control allocation strategy. This work con-
cludes that direct control allocation provides the smallest errors for it yields smaller deflections,
therefore control surfaces act more in their linear efficiency domain. As a recommendation the
authors suggest to be careful when pre-sizing control surfaces for BWB , for depending on the
allocation strategy chosen control surfaces efficiencies may be largely over-estimated.

1.2 Interactions of Aircraft Design and Control

While the previous section was specifically addressing the problem of BWB sizing and control,
this section deals with the more general topic of interactions between aircraft design and control.
Indeed adding control laws to enhance natural handling qualities of the airplane arises immediately
when designing a BWB . However also for conventional airplanes gains are expected of reducing
the aircraft natural stability, mostly by downsizing horizontal and vertical stabilizer. However
design of empennages and control surfaces then becomes challenging, because it requires taking
into account control laws already in the conceptual design phase. Section 1.2.1 details expected
gains of incorporating control laws early in the design phase of an aircraft. Then two concurrent
methods for designing simultaneously the airframe and control laws are examined: an MDO -based
approach is presented in section 1.2.2, and a control-based approach is explained in section 1.2.3.

1.2.1 Research in Control-Configured Vehicle

Concorde supersonic aircraft was the first commercial aircraft to incorporate Fly-by-Wire (FbW)
for control surfaces actuation [Favre, 1994]. On Concorde, orders sent to the control surfaces
were mostly proportional to the pilot column displacement, therefore reproducing the behavior of
mechanically actuated surfaces, with an additional pitch damper to increase longitudinal damping.
Yet FbW already revealed potentially very powerful to enhance natural handling qualities, because
of the possibility of including control laws for stability augmentation, flight envelope protection,
of structural loads alleviation. A research field therefore emerged in the seventies dedicated to
the exploration of potential airframe improvements by incorporating control laws effects at design
1.2. Interactions of Aircraft Design and Control 25

level: this field was named research in Control-Configured Vehicle (CCV).

[Abzug and Larrabee, 2005] provides an historical overview of stability and control technologies
incorporated into airplanes. According to the authors, the first artificially augmented airplane was
the Boeing B-47 bomber. Flight tests pilots discovered that yaw damping in low-speed approach
was too low, particularly the coupling with roll motion was found unacceptable. As a consequence
Boeing engineers decided to artificially damp the movement by feeding back a rate gyro to the
airplane rudder. While modern control theory was investigated in the design, it was reported that
the yaw damper gain was manually set in flight test by the chief test pilot. A noteworthy anecdote
is reported concerning the reaction of some eminent aircraft design professors to this artificial
solution of an aerodynamic issue: a CalTech professor, followed by a MIT professor, said: “If the
B-47 had been designed properly, it would not have needed electronic stability augmentation.” It
was also reported that the YB-49 yaw damper was designed, manufactured, and installed, and
flight-tested in five weeks.

Once overcome some previously mentioned reluctance about artificially augmented airplanes,
significant improvements of the airframe were expected with this approach. Still in [Abzug and
Larrabee, 2005], following gains are expected: “ maximum lift coefficient increases of 25% and trim
drag decreases of 20% can be obtained with a longitudinally unstable tailless design as compared
with its stable counterpart.” In [Ashkenas and Klyde, 1989], 10% range improvement of Relaxed
Static Stability (RSS) versus conventional configuration can be achieved, primarily because of
Czmax improvement. This in turn implies a wing area decrease for a tailless aircraft configuration,
with associated weight gains, and a tail area decrease for a tail-aft configuration. This is in
accordance with statements by [Roskam, 1995]: what drives airplanes designs towards artificial
stability is savings in tail areas and weight as well as savings in drag. Also in [Hartmann et al.,
1976], it is emphasized that flying at aft CG and therefore relaxing stability requirements implies
reducing the elevator down-lift, even possibly an uploaded tail. Trim drag is then significantly
reduced.

More precisely, following Control-Configured Vehicle (CCV) aspects were investigated in


literature:

1. Relaxed Static Stability (RSS): This involves reducing the aircraft static margin so that
reduced horizontal tail area and tail length are obtained. Expected benefits are improvements
of the wetted area and the horizontal tail weight, at the expense of a SAS .

2. Maneuver Loads Alleviation (MLA): This system aims at reducing the load on the wing,
particularly the root bending moment, during an aircraft maneuver by deflecting differentially
the control surfaces. See for instance [Gaulocher et al., 2007].

3. Gust Loads Alleviation (GLA): This system aims at reducing loads on the wing when facing
a gust. Weight savings on the wing structure are expected.

Our focus from now on will mainly be set on RSS concepts. In [Walker, 1976] a study is
performed in order to derive a CCV from a conventional tanker design with the same mission
26 Chapter 1. State-of-the-Art

requirements. In this study lateral and longitudinal SASs are designed in order to address RSS
challenges. Control laws are synthesized so that handling qualities satisfy [Chalk, 1969] military
specifications concerning poles characteristics. Output of the study is visible on figure 1.7.

Figure 1.7: Geometry of baseline and CCV tanker. Source: [Walker, 1976].

Among main conclusions, some are specifically addressing our problem:

1. Significant reductions in Gross Weight (GW) , Operating Weight Empty (OWE) and cost
are the major benefits resulting from the application of CCV concepts to transport type
airplanes.

2. Of all the CCV concepts, RSS has the most extensive impact on the airplane configuration
arrangement design and produces the largest reductions in weight and drag.

3. New handling quality criteria are needed because a demarcation between the short period
and phugoid modes is lacking at some flight conditions for the RSS airframe.

4. Although active controls were included in the preliminary design stage, the preliminary design
process for the CCV is standard in that, first, the airframe is statically designed after which
active control systems are designed.

Last remark is of particular importance: ”standard” preliminary design process is opposed to


“integrated” design process presented in section 1.2.2 and section 1.2.3, in which active control
systems are designed together with the airframe.

Another comprehensive study was conducted in the same period at the Cornell Aeronautical
Laboratory [Rynaski and Weingarten, 1972]. Starting from a baseline configuration of a T-33
aircraft, a direct optimization of tail length, tail area and control surface deflections required to
obtain similar handling qualities of initial aircraft and CCV is performed. Once again conclusions
of this study are noteworthy:
1.2. Interactions of Aircraft Design and Control 27

1. RSS , Maneuver Load Control, and good flying qualities can be made to be compatible if
adequate numbers of independent force/moment producing devices with adequate effective-
ness and power are provided.4

2. Because of the geometry and surface configurations are generally fixed, the application of
CCV concepts after the fact, i.e, on a presently existing aircraft, will yield only limited
success. CCV concepts, to be more effective, should be incorporated into the preliminary
design stages of a new airplane.

3. The present state-of-the-art 5 of feedback control allows for augmentation of an extremely


wide variety of bare airframe characteristics and therefore, geometrical shapes of the air-
frames. Stability constraints, such as static margins, have little or no importance if sufficient
control effectiveness and power are available to provide for good flying qualities and maneu-
verability.

A fundamental notion arises in the last point: there is no theoretical limit to the controllability
of an unstable aircraft, except the control power, effectors limits (bandwidth, saturations...), and
failure cases 6 .

Another study of interest is from [Anderson and Mason, 1996]. Besides presenting most
promising results in CCV design and historical perspective, they develop a method based on
fuzzy-logic to evaluate a control law complexity. The underlying idea is that a weighting function
of the control law complexity is needed to perform an aircraft design optimization which includes
control laws design. Otherwise the optimizer would necessarily converge towards a very high-order
control law and a highly unstable aircraft, a solution which would be practically infeasible.

More recently, a method for evaluating the gains achievable by relaxing stability is developed
in [Feuersänger, 2007].

1.2.2 Multidisciplinary Optimization Approach

As developed in the previous section, significant gains are expected from incorporating control
laws early in the aircraft design phase, in order to design a naturally unstable aircraft. However
the problem of geometrically designing the aircraft, which is for instance documented in [Roskam,
1985b] [Roskam, 1985a] [Roskam, 1985c], becomes coupled with the problem of designing control
laws, that is in a nutshell how to compute gains function of the aircraft dynamics and flight point to
ensure a safe aircraft behavior7 . This problem is in itself an active field of research, with specifically
developed routines and algorithms. Two concurrent approaches for integrating these two different
worlds therefore exist: the first one, presented in this section, consists in incorporating control
4
This issue is related to control allocation for redundant control effectors, and is treated in section 1.1.3.3.
5
In 1976.
6
However beyond a certain limit there is also no more gain to fly unstable
7
Strictly speaking the problem of controlling an aircraft is of course broader than that, and may be generally
defined as determining control surfaces behavior for given pilot and external inputs, to ensure a safe flight. This
task is not restricted to control gains computing, but also flight envelope protection design, modes switch, failure
cases handling...
28 Chapter 1. State-of-the-Art

laws design into a Multidisciplinary Optimization loop. The second one, developed in next section,
consists in taking advantage of recently developed control laws optimization tools to incorporate
some physical design together in the process.

A first study of interest concerning integration of FCS in an aircraft sizing and optimization
loop can be found in [Perez et al., 2006]. An MDO capability for aircraft sizing and optimization is
developed, that incorporates “traditional” aircraft design disciplines such as weight, aerodynamics,
propulsion and performance, but also a Stability & Control (S&C) module, which is the novelty
of this work. A conventional tail-aft aircraft is geometrically parameterized. Constraints are
set on geometrical variables, on TLARs such as takeoff and landing field length, and also on
flight dynamics and control by constraining control surfaces utilization, open- and closed-loop
longitudinal and lateral poles locations. Conventional control effectors, namely elevator, ailerons
and rudder are used. Two concurrent optimizations are evaluated: the first one does not include
control laws gains as design variables, therefore handling qualities must be achieved by the natural
aircraft. The second optimization includes control laws gains as variables, therefore leading to a
difference between open- and closed-loop poles locations. The design goal is to maximize the air
range for fixed fuel and payload weights. Collaborative Optimization (CO) was selected among
other MDO strategies – a comprehensive review on MDO architectures can be found in [Ruben
Perez et al., 2004] and [Martins and Lambe, 2013]. Comparison of the two optimized design shows
an improvement of 510N M air range for the CCV with respect to the “traditional” design. Main
geometrical differences involve horizontal and vertical tail sizes, implying significant gains in wetted
area. Open-loop handling qualities are then degraded compared to the baseline aircraft, but flying
characteristics are recovered by the inclusion of control laws. A global augmentation of control
power is observed, but within pre-defined limits. It must also be noted that the optimized design
remains stable, though with a reduced static margin compared to the conventional airplane.

Figure 1.8: Comparison of optimized designs, without and with flight dynamics and control con-
straints. Source [Perez et al., 2006]
1.2. Interactions of Aircraft Design and Control 29

In [Sahasrabudhe et al., 1997] a combined optimization of a helicopter rotor and flight control
system is performed. More precisely a sequential optimization, ie rotor design followed by control
design optimization for a fixed rotor, is compared to an integrated design of rotor and control
system. The optimizer is a Feasible Sequential Quadratic Programming (FSQP) , the objective
being to minimize control activity. Constraints include frequential, pole placement and temporal
constraints. While both designs satisfy all constraints, the integrated optimization induces lower
control effort requirements. As a conclusion of the study, it appears that rotor and control system
designs are closely coupled, therefore integrating both designs into a single optimization turns out
highly relevant.

More recently [Beaverstock et al., 2012] also integrated a flight dynamics and control module
into an aircraft design optimization process. Test-case on a Boeing 747-100 is analyzed: starting
from the initial design, the HTP area was reduced by 5% while longitudinal state-feedback α, q
control law was synthesized using eigenvalue assignment technique in order to retrieve the initial
aircraft dynamics. Unsurprisingly reducing HTP area leads to skin friction drag and trim drag
decrease, at the expense of reduced static stability which is recovered by stabilizing control laws.

Also genetic algorithms were tried on the integrated design and control problem in [Zhang
and Lum, 2007]. Aerodynamic model of a P-51 Mustang is computed using a light CFD panels
methods. An LQ regulator is computed for each design. The objective is a combined function
of the LQ cost function, accounting for control energy, and control surface size. Constraints
are set on open-loop poles location, and maximum trim deflection. The main interest of using
genetic algorithms is the ability to work with a multi-objective problem, and thus to provide a
Pareto front of different optimal solutions.

To conclude this part, let us remark that numerous studies in literature investigate MDO for
aircraft design incorporating disciplines such as performance, aerodynamics, weight and propulsion
– see for instance [Piperni et al., 2013]– however very few include a S&C module, taking into
account control laws, despite potential gains regularly forecast. In our view, this is mostly due
to the fact that, even though efficient procedures exist for control laws design, they are not yet
systematic enough to be embedded into an optimization procedure. Moreover expressing stability
and control requirements as optimization constraints is not straightforward.

1.2.3 Control-Based Approach

In the control community, the integrated design and control problem is often referred to as “plant-
controller optimization” problem. Indeed, historically control engineers are involved relatively late
in the design phase, when the airframe design is mostly frozen: a fixed plant is handed over to
them, generally based on engineering knowledge. Should control designers discover any deficiency
regarding the plant dynamics or its controllability during the control design phase, a costly come
back to the plan design phase would be needed. The idea of no more working on a fixed plant,
taking advantage of the advent of optimization techniques for control, then emerged.

From a theoretical point of view, [Fathy et al., 2001] mathematically studied the couplings
between plant and controller optimization problems. It is shown that even in the case of a
30 Chapter 1. State-of-the-Art

linearized state-space representation with state and control matrices, A and B respectively,
linear with respect to the plant variables, the problem becomes bilinear, non-convex. Four different
optimization strategies are then investigated:

1. sequential: this corresponds to the classical procedure of designing the plant first and then
the controller.

2. iterative: starting from an initial plant/controller design, a first plant optimization design
aims at improving plant design without compromising controller performance, and then the
controller design is optimized again without affecting plant performance.

3. nested: an outer loop optimizes the plant design, and for each plant design an optimal
controller is computed.

4. simultaneous: plant and controller designs are carried out simultaneously.

It is shown that iterative and sequential strategies are not guaranteed to converge towards
a global optimum, because plant and controller optimality conditions may be different. Also in
this study no particular structure is assumed, and state-feedback is assumed available. What the
results would become with constraints on controller structure or with an observer is left for future
work by the authors.

The plant/controller optimization problem was then applied in a variety of fields, such as
robotics [Ravichandran et al., 2006], structure [Velni et al., 2009], or chemistry [Ricardez, 2008]
[Ricardez-Sandoval et al., 2009] [Bansal et al., 2000].

In the aeronautic field, all studies to our knowledge focus on the HTP sizing and longitudinal
control law integrated design. Concerning problem formulation, [Hallberg, 1997] adresses the
following concern: how to capture the “size” of the control law? Indeed, if no constraint or
objective is set on the control law, then it may become arbitrary fast and powerful. The tradeoff
is expressed by the author as follows: "Since the degree of control of the longitudinal dynamics
depends on how fast and far the longitudinal control surface(s) can be moved by the control
actuator(s), a natural metric that captures the size of the automatic flight control system is
the maximum actuator rate. The design trade-off naturally includes consideration of actuator
performance. For instance, it may be more cost-effective to incorporate faster, generally larger
and more expensive, actuators rather than pay the drag penalty associated with a larger horizontal
tail." Limiting control law efficiency by constraining the actuators rate is an idea that we kept for
our study (see chapter 7). Also the perturbation used to evaluate control surface rate is an initial
perturbation on the angle of attack α . Following problem formulation of [Fathy et al., 2001],
an iterative optimization is performed. As a consequence minimizing the actuator rate limit is a
convex optimization that can be solved exactly using LMI framework.

This LMI framework was also applied for plant/controller optimization with aeronautical
application in [Niewoehner and Kaminer, 1996]. State-space representation is first formulated as
an affine approximation with respect to control surfaces size. Then an iterative optimization is run
using LMI : first control surface size – in this case HTP area – is optimized, then a controller
1.3. Conclusion 31

is optimized with fixed control surface size, and the algorithm loops until it converges. The main
interest of this article is that is expresses open- and closed-loop handling qualities constraints as
norms on H∞ transfer functions. With this formulation efficient H∞ -control algorithms can then
be used to optimize jointly the controller and the plant. Here LMI formulation, while powerful,
leads to full-order controller, i.e. no assumption can be made on the controller structure.

The LMI theory gained popularity among plant/controller design community, for it enables
convexifying the problem . This framework got for instance applied for integrated design and
control for underwater vehicle in [Silvestre et al., 1998]. More recently a similar approach was
developed in [Yang and Lum, 2003], and in [Liao et al., 2005a] the dependency of state-space
representation with respect to a unique plant variable is formulated using Linear Fractional
Transformation (LFT) framework. The case of two plant variables is addressed in [Liao et al.,
2005b], with a mixed H2 /H∞ formulation. The objective function is a weighting function of the
two plant parameters, the objective being to minimize control surfaces sizes.

Recently H∞ control theory gained a renewed interest thanks to the introduction of algo-
rithms for structured H∞ synthesis [Gahinet and Apkarian, 2011b] [Burke et al., 2006] [Gahinet
and Apkarian, 2012]. Constraints on controller structure implied using nonsmooth optimization
techniques, that proved more suited to the industry needs – for industrials often wish to keep a
structure of control laws that comes from a wide physical knowledge of the system to control
–, but also faster, less conservative, and better- conditioned than LMI techniques [Simon and
Wertz, 2011]. For those reasons [Alazard et al., 2013] applied non-smooth optimization techniques
for structured synthesis to a plant/controller optimization, for a satellite controller and avionics
design. In this paper plant/controller optimization is named “co-design”. A time delay accounting
for the avionics quality is optimized together with the control law synthesis. The expected trade
is that a worse avionics, symbolized by a large delay value, could be compensated by a more
complex controller. A cost function, built so that it penalizes short transmission delays, is set up.
Then the synthesis of both the controller and the avionics delay is performed with a structured
H∞ synthesis routine. This study indeed demonstrates possible trades between control law and
avionics complexity.

1.3 Conclusion

In this literature review, it was demonstrated that BWB configuration is very promising due to
high gains in wetted area, however severe stability and control challenges are yet to overcome.
This issue, while often pointed out in several studies, has not yet been addressed extensively
to our knowledge. More precisely the trade between instability and aerodynamic performance is
still an open question. Concerning control surfaces sizing, closed-loop design of these multi-axes
controls remains to be treated. In the aeronautical research literature, much work has been done
concerning the design of relaxed-stability vehicles, however always for conventional tail-aft design.
Integrated design and control, also called plant/controller optimization, was studied in literature
using different techniques such as Multidisciplinary Optimization or control approach. Most of the
time only longitudinal case is addressed, only sizing of the elevator or HTP area is addressed, and
32 Chapter 1. State-of-the-Art

approximations of aircraft stability derivatives dependency with respect to these physical variables
are limited to linear or affine approximations, so they lack precision. Finally comprehensive control
law structure based on industrial knowledge is not yet taken into account.
Part II

METHODS AND TOOLS

33
Chapter 2

Flight Dynamics Models

Contents
2.1 Flight Mechanics Equations . . . . . . . . . . . . . . . . . . . . . . . . 35
2.1.1 Coordinates Systems Definition . . . . . . . . . . . . . . . . . . . . 35
2.1.2 Fundamental Kinematics Equation . . . . . . . . . . . . . . . . . . . 39
2.1.3 Fundamental Mechanics Principle applied to the Airborne Aircraft . . 39
2.1.4 Longitudinal Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.1.5 Lateral Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.1.6 Concluding Remarks on Flight Dynamics Model . . . . . . . . . . . . 49
2.2 Blended Wing-Body Geometrical Model . . . . . . . . . . . . . . . . . 50
2.3 Description of Aircraft Flight Envelope . . . . . . . . . . . . . . . . . . 52
2.4 Open Loop Modes Analysis . . . . . . . . . . . . . . . . . . . . . . . . 53
2.4.1 Longitudinal Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.4.2 Lateral Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.5 Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

This “Methods and Tools” part is dedicated to describing the analysis tools that we will use
in the Scientific Contribution part. In this chapter the flight mechanics model of the BWB is
described and analyzed. More precisely flight mechanics equations are set in the general case,
then the geometrical and flight mechanics models used in our study are described.

2.1 Flight Mechanics Equations

In this section, the flight mechanics models are presented and analyzed. It is based on a six degrees
of freedom dynamic model. First references frames useful for equations writing are defined, that
is earth, aerodynamic and aircraft body reference systems respectively. Then forces and moments
that act on the aircraft are detailed and modeled, and a linearization of the equations is performed
in order to obtain a state-space representation suitable for analysis.

2.1.1 Coordinates Systems Definition

Three fundamental coordinate systems need to be defined in order to properly study and quantify
the aircraft motion. These coordinate systems are respectively:

• The earth reference system R0

35
36 Chapter 2. Flight Dynamics Models

• The aircraft body reference system Rb

• The aerodynamic reference system Ra

More details of each coordinate system, and how to express coordinates from one system to
another, are given hereafter.For a comprehensive theory on flight mechanics equations, the reader
may refer to [Taquin, 2009] [Boiffier, 1998].

2.1.1.1 Earth Reference System

The first coordinate system of interest is the earth reference frame R0 . Its origin is an arbitrary
point linked to the earth surface O, and it is composed of three direct orthonormal vectors
(x0 , y0 , z0 ), so it is noted R0 (O, x0 , y0 , z0 ). Vector Oz0 is vertical, oriented downwards, parallel
to the gravity vector. Ox0 y0 is the earth horizontal plane, and Ox0 is an arbitrary reference, for
instance magnetic or geographic North.

Assuming the hypotheses of a flat and fixed earth, which are valid for atmospheric subsonic
flight, this reference system becomes inertial. Then fundamental mechanics equations will be
expressed with respect to this inertial frame. This coordinates system is represented on figure 2.1.

2.1.1.2 Aircraft Body Reference System

A second reference system is the airplane body reference system Rb , with origin G the aircraft
center of gravity, and noted Rb (G, xb , yb , zb ). Vector Gxb is the fuselage horizontal reference,
oriented positive forward. Plane Gxb zb is the aircraft plane of symmetry, zb is oriented positive
downwards the aircraft. Finally yb completes the coordinates system so that (xb , yb , zb ) is positive,
leading to yb positive towards the right of the pilot.

This system is mainly important because all quantities measures inside an airplane, for example
in flight tests, are directly expressed in this system. For instance an accelerometer will measure
three-axes accelerations which are aircraft body-related (obviously because sensors are linked to
the airframe).

Rotation vector from Rb to R0 projected in Rb 1 is simply written as:


 
p
ΩRb /R0 = q  (2.1)
 

r Rb

A set of three angles allows transformation from earth reference frame R0 to aircraft body
frame Rb :
1
Let us recall that any vectorial relation can be projected in any frame: derivation with respect to a given frame
is independent from projection of the derived vector in any arbitrary frame.
2.1. Flight Mechanics Equations 37

Figure 2.1: Earth and Aircraft Body Reference Systems.

• Ψ is the azimuth angle between x0 and xΨ the projection of xb on the horizontal plane
Gx0 y0 .

• θ is the pitch attitude angle between xΨ and xb .

• φ is the bank angle, also the angle between yΨ the projection of yb on the horizontal plane
Gx0 y0 and yb .

All these angles are depicted on figure 2.1.

2.1.1.3 Aerodynamic Reference System

The last coordinates system that we will use in our study is the aerodynamic reference system Ra
. Once again its origin is G the aircraft center of gravity, and is noted Ra (G, xa , ya , za ). Vector
Gxa is directed along the air velocity vector Vair . Gza is perpendicular to Gxa and within the
aircraft plane of symmetry. Finally Gya defines the final direct triedral. This system is presented
on figure 2.2.

Two angles allow transformation from Ra to Rb :

• β is the sideslip angle between ya and yb .


38 Chapter 2. Flight Dynamics Models

Figure 2.2: Aerodynamic and Aircraft Body Reference System.

• α is the angle of attack between x2 the projection of xa on the aircraft plane of symmetry
and xb .

Also three angles are necessary to perform a transformation from R0 to Ra :

• γ is the flight path angle, defined as the angle between xa and the horizontal plane Gx0 y0 .

• χ is the aerodynamic azimuth, which is the angle between x0 and xχ the projection of xa
on the horizontal plane Gx0 y0 .

• µ is the aerodynamic bank angle, measure of the angle between ya and the horizontal plane
Gx0 y0 .

These angles are visible on figure 2.3. Among them only one is of practical use for our study: the
flight path angle γ .

The main interest of the aerodynamic reference system Ra is that aerodynamic forces are
expressed in a convenient form in this system.
2.1. Flight Mechanics Equations 39

Figure 2.3: Aerodynamic and Earth Reference Systems.

2.1.2 Fundamental Kinematics Equation

In order to efficiently compute a vector temporal derivative with respect to an arbitrary reference
frame R0 , following fundamental kinematics relationship is used:

dX dX
= + ΩR1 /R0 ∧ X (2.2)
dt R0 dt R1

Indeed vector X may be more easily derivated in the reference frame R1 , however cross-
product of rotation vector of R1 with respect to R0 and X must be added in order to get the
derivative in reference frame R0 .

2.1.3 Fundamental Mechanics Principle applied to the Airborne Aircraft

General equations of flight dynamics are directly derived from the fundamental mechanics principle,
that states that the inertial acceleration of a body multiplied by its mass is equal to the sum of the
external forces. In our study we only consider the aircraft as airborne, that is no ground reaction
is considered. As a consequence the only forces applied to the aircraft are its weight, aerodynamic
40 Chapter 2. Flight Dynamics Models

and propulsive forces. Hence the fundamental mechanics principle reads:

mV˙g = mg + Faero + Fthrust (2.3)


G G G G
J Ω̇ = Maero + Mthrust (2.4)

In equations (2.3) and (2.4) m refers to the aircraft mass that is supposed constant2 , Vg is
the aircraft ground speed, that is the speed of G its center of gravity with respect to the inertial
frame R0 , g is the gravity vector, Faero and Fthrust represent the aerodynamic and propulsive
forces acting on the aircraft respectively. J G is the aircraft inertia matrix with respect to its CG
, ΩG is the aircraft rotation with respect to R0 , Maero G G
and Mthrust are the aerodynamic and
propulsive moments with respect to the CG respectively. Moments equation being considered
with respect to the CG , obviously no moment due to weight needs to be considered. Would
equation (2.4) be considered with respect to another point, moment due to weight should be
added.

An alternative formulation of equation (2.3) can be written by introducing the load factor n
:
mgn = Faero + Fthrust (2.5)
n is the apparent weight, equal to the sum of non-inertial forces, a-dimensioned by weight mg.

2.1.3.1 Models for Forces and Moments

In this section forces and moments introduced in equation (2.3) are developed.

Ground Speed: Equations here are written considering the assumption of calm wind.3 Therefore
we have:

Vg = Vair (2.6)
= Vair xa (2.7)

by definition of xa .

Weight Force: By definition of R0 , g is along z0 :

g = g.z0 (2.8)
2
Rigorously the aircraft mass, as well as its CG location, are varying during flight because of fuel consumption.
The constant approximation is based on the rationale that variations are slow enough not to affect short-term
dynamics, typically with an order of magnitude of a few seconds, that are our focus.
3
In general case following fundamental equation yields: Vg = Vw + Vair , where Vw is the wind speed.
2.1. Flight Mechanics Equations 41

Load Factor n may be projected in Ra with components written as follows:

nx
n= ny (2.9)
Ra
nz

Aerodynamic Force: Faero may be indifferently projected in Ra or Rb .

−Cx
1
Faero = ρVair 2 S Cya (2.10)
2
Ra
−Cz
−Ca
1
= ρVair 2 S Cy (2.11)
2
Rb
−Cn

From a practical point of view only Cx , Cz and Cy coefficients are used: as developed
hereafter, longitudinal equations are usually written in the aerodynamic system Ra and lateral
equations are written in the aircraft system Rb . Minus coefficients in equation (2.10) originate
from a sign convention: lift and drag coefficients Cx and Cz are considered positive.

Propulsive Force: Thrust is assumed parallel to the aircraft fuselage 4 , therefore Fthrust is
along xb :
Fthrust = Fthrust xb (2.12)

Aircraft Inertia: J G is expressed in the aircraft body reference system Rb . Inertia radius
notations are used, giving:
 
Rxx 0 Rxz
JG = m 0 Ryy 0  (2.13)
 

Rxz 0 Rzz R
b

Cross-inertia Rxy and Ryz terms are canceled under the hypothesis of Gxb zb being a symmetry
plane for the aircraft.

4
Strictly speaking a small setting angle always exist between thrust direction and the aircraft fuselage. Typical
order of magnitude on commercial aircraft is one degree, it is not considered in our study.
42 Chapter 2. Flight Dynamics Models

Aircraft Rotation Vector: rotation of the aircraft with respect to R0 ΩRb /R0 is expressed
in the aircraft reference frame Rb as:
 
p
ΩRb /R0 = q  (2.14)
 

r R
b

G
Aerodynamic Moments: Maero is written in the aerodynamic reference system Ra :

ClG
G 1
Maero = ρVair 2 Sl CmG (2.15)
2
Ra
CnG

Propulsive Moments: All engines are assumed operative, producing an equal thrust. For this
reason only thrust pitching moment is considered, giving:

G G
Mthrust = Mthrust .yb (2.16)

On the BWB engines are located above the CG , so thrust creates a nose-down pitching
G
moment, and Mthrust < 0.

2.1.4 Longitudinal Model

Previous equations are valid in the general three-axes case. For stability and control analysis, as
well as control laws design, it is relevant to decouple those equations into two sets of longitudinal
and lateral equations respectively, based on standard hypotheses. This section is devoted to
formalizing longitudinal equations, while section 2.1.5 will be dedicated to similar work on lateral
axes.

2.1.4.1 Hypotheses

The basic hypothesis for considering longitudinal flight is that all forces acting on the aircraft lie
within the aircraft symmetry plane (Gxb zb ); then rotation vectors and moments are normal to
this plane. From this it results:

• Weight force vector lies within the aircraft symmetry plane, meaning z0 is in the (Gxb zb ).
From figure 2.1 it results that bank angle φ equals zero: wings are in the horizontal plane.

• Aerodynamic forces also lie within the aircraft symmetry plane: according to figure 2.2, it
implies that sideslip β and aerodynamic bank angle µ are zero.
2.1. Flight Mechanics Equations 43

• Rotation vector being normal to the aircraft symmetry plane implies roll and yaw rates p
and r equal zero respectively. Only pitch rate q is not zero.

2.1.4.2 Kinematics Equations

For a comprehensive treatment of kinematics equations in the general case, the reader may refer to
[Taquin, 2009] [Boiffier, 1998]. For sake of brevity only particular cases are developed: longitudinal
kinematics is examined here, and lateral kinematics are developed in section 2.1.5.2.

From longitudinal flight hypotheses following vectors relationship holds: y0 = ya = yb . As


a consequence following angular relationship holds:

θ =α+γ (2.17)

Another consequence of assuming longitudinal flight is following relationship between pitch


rate and pitch attitude:
θ̇ = q (2.18)

Finally using equation (2.2) in the simplified case of longitudinal flight, and assuming no wind
as already stated in section 2.1.3.1, derivative of ground speed Vg with respect to earth frame
R0 gives:
dVg
= V̇air xa − Vair za (2.19)
dt R0

2.1.4.3 Fundamental Principle of Dynamics for Longitudinal Case

Considering previous assumptions, equation (2.3) projected on the aerodynamic frame Ra gives:

1
xa : mV̇air = − ρVair 2 SCx − mg sin γ + Fthrust cos α (2.20)
2
1
za : −mVair γ̇ = − ρVair 2 SCz + mg cos γ − Fthrust sin α (2.21)
2
1
ya : mRyy q̇ = ρVair 2 SlCmG (2.22)
2

Moreover equations (2.20) and (2.21) can be written using load factor components introduced
in equation (2.9):

1
xa : mgnx = Fthrust cos α − ρVair 2 SCx (2.23)
2
1 2
za : mgnz = ρVair SCz + Fthrust sin α (2.24)
2
44 Chapter 2. Flight Dynamics Models

2.1.4.4 Equations Linearization

Equations linearization around an equilibrium allows using linear analysis tools for dynamics analy-
sis, and is therefore valuable. Aerodynamic models and linear approximations leading to a suitable
state-space representation for analysis are detailed here.

Aerodynamic Models Lift coefficient is modeled considering major contributions of the angle
of attack α , pitch rate q and control surfaces deflections δmi .

l
Cz = Czα (α − α0 ) + Czq q + Czδmi δmi (2.25)
Vair
Czα is the lift gradient versus angle of attack, α0 is zero-lift angle of attack, Czq is the lift
gradient versus pitch rate, and Czδmi physically corresponds to direct lift induced by deflecting
i-th trailing edge elevon. Such effect is present on conventional aircraft – downwards deflection
of the elevator induces a direct lift increment – but is particularly at stake for BWB dynamics,
for the elevons directly modify main lifting surface camber.

As developed in section 2.2, BWB configuration studied here features ten indepen-
dently moving control surfaces on the trailing edge. Contributions of all control surfaces
to aerodynamics coefficient sum up, yet it was chosen not to display the summation sign
in the equations for sake of clarity. Thus from now on δmi should be understood as
P10
δmi , ∀i = 1 . . . 10. Similarly CmG
δmi should be understood as
G
i=1 Cmδmi δmi .

Drag coefficient is modeled by including an induced drag model:

Cx = Cx0 + ki Cz 2 (2.26)

Performance analysis would require a more detailed drag model including compressibility, wave
and friction drag for instance; however our study focuses on stability and control, for which model
described in equation (2.20) is considered sufficient.

Pitching moment coefficient is modeled as follows:

XF − XCG l
CmG = Cm0 − Czα (α − α0 ) + CmG
q q + CmG
δmi δmi (2.27)
| l
{z } Vair
ms

Cm0 is zero-lift pitching moment, ms = XF −Xl


CG
is the aircraft static margin, representing
relative position of the aerodynamic center XF and center of gravity XCG . CG being
classically expressed as a percentage of the Aerodynamic Mean Chord (AMC) and oriented
positive backwards, a positive static margin ms > 0 is equivalent to XCG lying forward the
aerodynamic center XF , and in that case the aircraft is said statically stable. Indeed starting
from an equilibrated position CmG = 0, a positive angle of attack increment ∆α > 0 induces a
pitch down moment CmG (∆α) = −ms Czα ∆α < 0 which makes the aircraft move back to its
2.1. Flight Mechanics Equations 45

initial equilibrium.

Longitudinal Equilibrium Linearization is performed around an equilibrium point, that is a


flight point where states variables are constant. In this study altitude is not considered as a state
variable, so equilibrium at constant flight path angle can be considered. Cancellation of derivative
terms in equations (2.20) to (2.22) then leads to equilibrium equations:

1
xa : Fthruste cos αe + mg sin γe = ρVe 2 SCxe (2.28)
2
1
za : mg cos γe + Fthruste sin αe = ρVe 2 SCze (2.29)
2
ya : CmG e =0 (2.30)

Fthruste , αe , γe , Ve , Cxe , Cze and CmG e refer to equilibrium thrust, angle of attack,
flight path angle, airspeed, drag and lift coefficient respectively.

Linearized Equations For linearization of equations (2.20) to (2.22) around equilibrium points
defined in equations (2.28) to (2.30), some hypotheses are made:

• On xa axis, αe is assumed small, so cos αe ≈ 1 and equation (2.28) becomes:

1
Fthruste = ρVe 2 SCxe − mg sin γe (2.31)
2

• On za axis:

– Same hypothesis on αe being small leads to neglecting Fthruste sin αe term.


– Equilibrium flight path angle γe is also assumed small, so cos γe ≈ 1 .
– Hence equation (2.29) becomes:

1
mg = ρVe 2 SCze (2.32)
2

Finally linearizing equations (2.20) to (2.22) gives:

ρVe S 2g ρSl ρVe S


δ γ̇ = Czα δα + 2 δV + Czq q + Czδmi ∆δmi (2.33)
2m Ve 2m 2m
1 ∂F 2gki Czq l
 
δ V̇ = − ρVe SCx δV − 2gki Czα δα − q...
m ∂V Ve
1 ∂F
· · · − 2gki Czδmi ∆δmi − gδγ + (2.34)
 m ∂∆x
ρVe S 2g ρSl ρVe S

δ α̇ = − Czα δα − 2 δV + 1 − Czq q − Czδmi ∆δmi (2.35)
2m Ve 2m 2m
ρVe 2 Sl ρVe Sl2 ρVe 2 Sl
q̇ = CmGα δα + Cm G
q q + CmG δmi ∆δmi (2.36)
2mRyy 2mRyy 2mRyy
46 Chapter 2. Flight Dynamics Models

where δγ = γ − γe , δα = α − αe , δV = Vair − Ve , ∆δmi = δmi − δmie and ∆δx = δx − δxe are


differences between variables and their initial equilibrium values.

2.1.4.5 State-Space Representation

State-space representation of linearized equations enables applying useful tools of linear systems
theory to understand the aircraft dynamics, such as eigenvalues analysis, as well as controllability
and observability analyses. General Linear Time-Invariant (LTI) state-space representation is:
(
Ẋ = AX + BU (2.37)
Y = CX + DU (2.38)

In our study longitudinal states are chosen as follows:


 
δVair
 δα 
Xa =  (2.39)
 
 q 

δθ

Using fundamental equation for longitudinal flight 2.18 together with previously developed
linearized equations, following state-space representation holds:

      
δ V̇ xV xα xq xθ δV xδx xδmi !
 δ α̇  −z −zα 1 − zq 0   δα   0 −zδmi 
 ∆δx
   
   V
 =  + (2.40)
 q̇   0 mα mq 0  q   0 mδmi  ∆δmi

δ θ̇ 0 0 1 0 δθ 0 0 | {z }
Ua
| {z } | {z } | {z }
Aa Xa Ba

Expression of matrices coefficients is given in appendix A.

In aircraft control it is usually assumed that all states variables are measurable. However
for longitudinal control another measure is added to states measurements: vertical load factor
variation δnz . Identification of equations (2.21) and (2.24) gives:

Vair
δnz = δ γ̇ (2.41)
g
2.1. Flight Mechanics Equations 47

Consequently output vector Ya = (Xa δnz )T is obtained as follows:

···
 
1 0 0  
 .. ..  0
!  0 . .  . 
Xa    .. 
=  ... X +  ∆δmi (2.42)
 
δnz  1 0 a  0


| {z }  0 ··· 0 1

Ya Ve Ve Ve
zδmi
g zV g zα g zq 0 | {z }
| {z } Da
Ca

2.1.5 Lateral Model

Similarly to what was developed insection 2.1.4, in this section equations for lateral approximation
are developed. First hypotheses for lateral flight are established, then kinematics and dynamics
equations are set, and finally a state-space representation is derived from linear representation.

2.1.5.1 Hypotheses

Unlike longitudinal model, which is perfectly decoupled from lateral states provided adequate
hypotheses are fulfilled, lateral model cannot be completely decoupled from longitudinal states.
Yet a standard hypothesis for studying separately lateral states is to assume that longitudinal
variables Vair , α , θ , q are constant, that is piloted.

Then cross-axes Rxz inertia term will be neglected for sake of conciseness.

2.1.5.2 Kinematics Equations

Kinematics relations for lateral case are not as straightforward as for longitudinal flight. Applying
equation (2.2) to the aircraft lateral rotation gives:

φ̇ = p + (q sin φ + r cos φ) tan θ (2.43)

2.1.5.3 Fundamental Principle of Dynamics for Lateral Case

Unlike longitudinal case, lateral equations are more easily expressed in the airplane reference system
Rb . Applying equation (2.3) and projecting on yb axis for force equation and (xb , zb ) axes for
moments equations gives:
48 Chapter 2. Flight Dynamics Models

1
yb : m[Ve cos β β̇ + Ve cos β(r cos αe − p sin αe )] = ρVe 2 SCy + mg cos θ sin φ (2.44)
2
1
xb : m[Rxx ṗ + (Rzz − Ryy )qr] = ρVe 2 SlClG (2.45)
2
1
zb : m[Rzz ṙ + (Ryy − Rxx )pq] = ρVe 2 SlCnG (2.46)
2

2.1.5.4 Equations Linearization

Following a similar approach to what was developed for longitudinal case, lateral equations are
linearized for a latter modes analysis.

Aerodynamic Models Linear dependencies of aerodynamic coefficients with respect to states


and control variables are considered. More precisely, lateral force coefficient incorporates depen-
dencies with respect to sideslip, roll rate, yaw rate and rudder deflection:

l l
Cy = Cyβ β + Cyp p + Cyr r + Cyδn δn (2.47)
Ve Ve

Similarly aerodynamic yawing moment coefficient is written as:

l l
CnG = Cnβ β + Cnp p + Cnr r + Cnδn δn (2.48)
Ve Ve

Finally aerodynamic rolling moment coefficient is expressed as:

l l
ClG = Clβ β + Clp p + Clr r + Clδmi δmi (2.49)
Ve Ve
Remark 2.1
As it can be seen on figure 2.4, the Airbus BWB features two separate rudders, each one
mounted over an engine. For control purpose it was not considered necessary to distribute different
deflections on each rudder: therefore from now on a single equivalent rudder will be considered
for control, whose deflection is noted δn , and whose aerodynamic effectiveness is the sum of the
two separate rudders.

Linearized Equations Linearization is performed around cruise equilibrium points, at zero sideslip.
For this reasons following hypotheses are assumed:

• cos β ≈ 1 and sin β ≈ β.

• cos φ ≈ 1 and sin φ ≈ φ.


2.1. Flight Mechanics Equations 49

Linearization of equations (2.43) to (2.46) then gives:

ρVe S l l g
 
δ β̇ = p sin α − r cos α + Cyβ β + Cyp p + Cyr r + Cyδn ∆δn + cos θδφ
2m Ve Ve Ve
(2.50)
ρVe 2 Sl l l
 
ṙ = Cnβ β + Cnp p + Cnr r + Cnδn ∆δn (2.51)
2mRzz Ve Ve
2
ρVe Sl

l l

ṗ = Clβ β + Clp p + Clr r + Clδmi ∆δmi (2.52)
2mRxx Ve Ve
δ φ̇ = p + r tan θ (2.53)

2.1.5.5 State-Space Representation

Classically lateral states are chosen as follows:


 
β
r 
X=  (2.54)
 
p
φ

Following state-space representation results:

   g cos θ     
β̇ yβ yr − cos α yp + sin α Ve β 0 yδmi yδn !
 ṙ    r  0 n nδn 
  nβ nr np 0 δmi  ∆δmi
 = + (2.55)
  
 ṗ   lβ  p 0 lδmi lδn  δn
  
lr lp 0
φ̇ 0 tan θ 1 0 φ 0 0 0

Literal expression of matrices coefficients is given in appendix A.

2.1.6 Concluding Remarks on Flight Dynamics Model

The analytical model developed in this section aims at understanding the physical
principles of the aircraft dynamics, and the influence of different aerodynamic and inertia
parameters on the aircraft modes. Such analysis will be performed next. Later in
the study, control laws are performed on three-axes linear approximation of the BWB
for various flight points. Those state-space representations are not computed from
previously developed analytical expressions, rather they come from a finite-differences
approximation of a full non-linear model developed in [Saucez, 2013]. Therefore state-
space representations used for synthesis and simulations are richer than the models just
presented: some approximations and cross-coupling terms omitted here for sake of clarity
are implicitly taken into account in the rest of our study.
50 Chapter 2. Flight Dynamics Models

2.2 Blended Wing-Body Geometrical Model

δmL R
1 δm1

δmL
3 δmR
3
δmL
4 δmR
4

δmL δmL
2 δmR
2 δmR
5 5
δnL δnR

Figure 2.4: Planform view of the Airbus Blended Wing-Body.

This section briefly describes the BWB geometrical model used in this study. Data originate
from Airbus Future Projects Office (FPO) , which hosted our work. This BWB , whose planform
view is displayed on figure 2.4, has been studied for years as a potential candidate for future long
range program, and led to several publications [Saucez, 2013; Meheut et al., 2012; Saucez and
Boiffier, 2012].

In aircraft design vocabulary, design requirements are often designated “Top-Level Aircraft
Requirements”. Significant TLARs for this configuration are summarized in table 2.1.

TLAR
Design Range 7200N M
Passengers Capacity 470 pax in Mix-Class cabin arrange-
ment
Cruise Mach Number 0.85
Maximum Operating Speed 340kt
Span ≤ 80m
Cruise Altitude 35000f t ≤ h ≤ 43000f t
Take-Off Field Length (TOFL) ≤ 3000m at International Standard
Atmosphere (ISA) +15 and sea level
Approach Speed ≤ 150kt

Table 2.1: Top-Level Aircraft Requirement for Airbus Long-Range BWB .


2.2. Blended Wing-Body Geometrical Model 51

Major geometrical parameters describing the aircraft are gathered in table 2.2. Those figures
are only approximate since the exact aircraft model is confidential Airbus property. Yet presented
figures illustrate well the unusual dimensions of such aircraft. The span is similar to that of the
Airbus A380 for it is constrained by airports operations regulations. Then the aircraft length is
much shorter than that of a conventional aircraft. As a consequence longitudinal lever arms of
control surfaces are expected to be shorter than conventional elevators, and their longitudinal
effectiveness are expected to be reduced.

Reference area S 1200m2


Aerodynamic Mean Chord AMC 25m
Span b 80m
Root Chord 40m
Masses From 200t to 350t

Table 2.2: Geometry data of the configuration.

Elevon Min/Max Deflection Area


1 ±30˚ 11.8m2
2 ±30˚ 38.5m2
3 ±30˚ 17.3m2
4 ±30˚ 11.2m2
5 ±30˚ 6.5m2

Table 2.3: Control Surfaces Attributes.

Numerous redundant trailing edge control surfaces are a major peculiarity of BWB configu-
ration. On figure 2.4, 10 independent elevons are visible, denoted 1 to 5 from inboard to outboard
on each wing. Attributes for each control surface are summarized on table 2.3. Of particular
concern is the control surfaces area: for instance elevon 2 is 38.5m2 , which is approximately equal
to an A320 HTP area. However while the A320 HTP only moves for trim purpose at very
slow rates – typically less than 0.5˚/s–, BWB control surface is devoted to trim, maneuvering
and active stabilization. Therefore it requires higher deflection rates, in the order of magnitude
of ailerons or elevators i.e. 30 to 60˚/s, the exact deflection rate required being a desired output
of this study. Moving 40m2 at 60˚/s is likely to lead to high power requirements, justifying the
need for a preliminary assessment of stabilization and systems energy needs, which is the aim of
our work.

Significant control surface sizes are also clear when examining elevon hinges size. For instance
figure 2.5 displays a comparative view of the 2.30m hinge of elevon 2 and a 1.80m human. Even
though no detailed design of control surfaces geometry was performed from a structural point of
view prior to our work, the question of feasibility of such design for possibly rapidly moving control
surface is raised.
52 Chapter 2. Flight Dynamics Models

Figure 2.5: Illustration of significant hinge of elevon 2.

2.3 Description of Aircraft Flight Envelope

A proper aircraft sizing from stability and control perspective requires scanning the whole flight
envelope, in terms of speed, altitude, mass, CG and configurations. It was shown in previous
studies [Saucez, 2013] that this configuration does not call for any high-lift devices such as flaps
or slats; as a consequence only clean configuration was studied. Moreover the main purpose of our
study is to examine the impact of instability on controls and FCS sizing. Longitudinal instability
is a direct function of CG location at a given flight point, see equation (2.27). For this reason
only most backward CG position for each mass, given by weight & CG diagram, was considered.

Then four masses where examined: m = [200t, 250t, 300t, 350t]. Mach numbers are dis-
cretized from 0.25 to 0.85, with steps of 0.1, and for each Mach number a sweep is performed
on the airspeed Vair with steps of 20kts. Varying the airspeed for a given Mach number is
equivalent to varying the altitude according to the classical definition of Mach number:

Vair
M=p (2.56)
γ rair T
air

where T is linked to the altitude H through the hypothesis of International Standard Atmo-
sphere (ISA). Admissible altitudes are given by Velocity Maximum Operating (VMO) limitation
for low-altitude limit, and maximum of buffet lift coefficient Cz buffet and operational ceiling for
high-altitude limit. Buffetting is an instationary aerodynamic phenomenon that corresponds to
a high-altitude stall; for preliminary evaluation Cz buffet may be computed using knowledge on
Czmax value. Resulting flight envelope considered in our work is depicted on figure 2.6.
2.4. Open Loop Modes Analysis 53

Flight enveloppe
14000

12000

10000
Altitude (m)

8000

6000

4000
mass=200T
mass=250T
2000
mass=300T
mass=350T
0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Mach

Figure 2.6: Mach-Altitude flight envelope.

2.4 Open Loop Modes Analysis

The major purpose of state-space representations presented in sections 2.1.4.5 and 2.1.5.5 is to
provide tools from linear theory in order to derive information on the aircraft behavior, such as
eigenvalues analysis giving information on the aircraft modes. Section 2.4.1 is devoted to analyzing
longitudinal modes, which behave rather unusually on this aircraft, whereas section 2.4.2 treats
lateral dynamics.

2.4.1 Longitudinal Modes

Let us consider only longitudinal modes X = (δV δα q δθ)T , for which the state matrix is given in
equation (2.40). On classical tail-aft airplane two separate modes arise, both periodic and stable,
but of different time constants and acting on difference state variables:

• Short-Period mode is a fast converging mode, with a frequency ω0 ≈ 1rad/s. It mainly


acts on the angle of attack α and pitch rate q , hence classically this mode is studied by
isolating the subsystem (α, q) from longitudinal state-space representation.

• Phugoid Mode is a slow periodic motion which can be interpreted as an exchange of kinetic
and potential energy, so it mainly acts on the airspeed Vair and flight path angle γ .
54 Chapter 2. Flight Dynamics Models

A common hypothesis involves a constant angle-of-attack α during the movement, then


the subsystem (Vair , γ), or equivalently (Vair , θ), usually describes the phenomenon in a
satisfactory way.

However for the BWB , identification of those two modes is not straightforward. If they do exist,
their behavior is quite different from what was previously explained.

Longitudinal poles for H=3236m , mass=350T, M:0.35


0.15
0.996 0.993 0.986 0.965 0.86
Longitudinal Model
0.998 Reduced Model [V; 3]
0.1
Reduced Model [,; q]

0.999
Imaginary Axis (seconds -1)

0.05
1

2 1.75 1.5 1.25 1 0.75 0.5 0.25


0 1

1
-0.05

0.999

-0.1

0.998

0.996 0.993 0.986 0.965 0.86


-0.15
-2 -1.5 -1 -0.5 0 0.5 1
Real Axis (seconds -1)

Figure 2.7: Comparison of Longitudinal Poles with Reduced State-Space Representations [α, q]
and [V, θ].

Figure 2.7 displays the longitudinal poles on complex plane for a low Mach, heavy mass flight
point, together with poles coming from two second-order reduced state-space representations [α, q]
and [V, θ] respectively. From this figure we see that:

• [α, q] approximation is valid, relative error between actual and approximated poles being
below 0.3%. Yet poles location differs from “usual” short-period behavior: here two real
poles are visible, one lying in the right half-plane, that is unstable, with a frequency of
approximately 0.9 rad.s−1 . The topic of instability, which is a main concern in our study,
is specifically developed in next section 2.4.1.1.

• [V, θ] approximation is a bit less relevant, with a 12% relative error. “Phugoid” mode on this
aircraft is stable but with very low damping (ξ ≈ 0.03) and a frequency ω ≈ 0.12 rad.s−1 .
2.4. Open Loop Modes Analysis 55

2.4.1.1 Short-Period Mode

Let us analyze the origin of longitudinal instability. For that purpose let us express short-period
approximation [δα, q] using notations introduced in section 2.1.4.5:
! ! ! !
δ α̇ −zα 1 − zq δα −zδmi
= + ∆δmi (2.57)
q̇ mα mq q mδmi

Transitory dynamics is given by the eigenvalues of matrix A , which is equivalent to computing


roots of characteristic polynomial |sI − A|, where s is the Laplace variable and I is the identity
matrix of same size as A :

s + zα −(1 − zq )
|sI − A| =
−mα s − mq
|sI − A| = s2 + s(zα − mq ) − zα mq − mα (1 − zq ) (2.58)

As it was presented on figure 2.7, this aircraft features two real poles, that we call s1 and s2
respectively. Equation (2.58) is then identified with:

(s − s1 )(s − s2 ) = s2 − (s1 + s2 )s + s1 s2 (2.59)

Stability condition on equation (2.59) gives: s1 s2 > 0 5 , which translates into:

mα (1 − zq ) < −zα mq (2.60)

mq is linked to CmG q value, which is the pitch damping coefficient. This aerodynamic
coefficient is always negative; on a classical tail-aft airplane a positive pitch rate q > 0 induces
a downward motion of the HTP , which “sees” a higher local angle of attack than that of the
overall aircraft. As a consequence lift on the tail is increased, implying a pitch-down moment
opposing the pitch-up motion.

This CmG q effect is very small, and can be considered as negligible, on BWB for it lacks
any HTP . If CmG q , and also Czq effect which is small, are neglected, then equation (2.60)
becomes:

mα < 0 (2.61)
⇐⇒ CmG
α <0 (2.62)

which demonstrates the static stability condition already stated in section 2.1.4. Recalling that
CmG α is linked to XCG and XF relative positions:

XF − XCG
CmG
α =− Czα (2.63)
l
5
and also s1 + s2 < 0
56 Chapter 2. Flight Dynamics Models

necessary condition of XCG lying forward XF 6 for positive static stability is found again.

Yet if starting again from equation (2.60), Czq and CmG


q coefficients are no more neglected,
then stability condition reads:
−zα mq
mα < (2.64)
1 − zq
zq term being always small compared to 1, and mq < 0, it results that the aircraft may remain
stable even with a slightly positive CmG
α . More precisely equation (2.64) implies:

ρSl2 CmG
q
XCG < XF − ρSl
(2.65)
1− 2m Czq
| {z }
XFq

A new point XFq appears, called the maneuver point, which defines a new dynamic stability
criterion. XFq is located backward XF , meaning the CmG q effect allows remaining stable
with a CG beyond the aerodynamic center.

However while on a conventional airplane the difference between XF and XFq may be
significant due to the empennage damping effect, on the Airbus BWB the CmG q term is
negligible, implying the aerodynamic center XF and maneuver point XFq are almost the same.

Figure 2.8: Position of aerodynamic center XF with respect to CG in high and low speed.

On figure 2.8, relative position of XF and XCG is shown in high and low speed. Clearly
XF lies on the left of the Weight & Center of Gravity (W&CG) diagram, whereas stability
condition would require it lies on the right (abscissae are positive towards the rear of the aircraft).
6
Let us recall that XCG measurement is oriented positive toward the rear of the aircraft.
2.4. Open Loop Modes Analysis 57

Moreover compressibility effects tend to move XF backward, implying the most critical situation
from instability point of view is in low speed. Yet in low the control surfaces are least efficient
due to low dynamic pressure 21 ρVair 2 : these two effects are detrimental. It can also be noted that
for high speed cruise and at specific mass and foward CG configuration –namely huge payload
and low fuel–, the airplane gets stable.

The evolution of the aerodynamic center, that is level of instability, with respect to speed is
corroborated by figure 2.9. Longitudinal poles are plotted for different Mach numbers at a given
altitude and mass: only speed varies. It can be seen that the higher the speed, the less instability
is obtained.

Figure 2.9: Evolution of Longitudinal Poles with Speed.

2.4.1.2 Phugoid Mode

As already emphasized previously, whereas the short-period approximation truly represents two of
the four longitudinal modes, the phugoid approximation of explaining the two remaining modes by
considering only the [V, θ] subsystem is not fully satisfactory. A clue for this lack of representative-
ness is visible on figure 2.9. Whereas common phugoid explanation demonstrates a proportional
relationship between motion period and equilibrium speed Ve :
s
2
Tphugoid ≈ Ve (2.66)
g
58 Chapter 2. Flight Dynamics Models

on figure 2.9 on the contrary, “phugoid” frequency ωphugoid increases with speed. For those
reasons these two remaining poles should not be considered as a “phugoid” properly speaking.

Yet from a handling qualities perspective this mode is of little concern: while poorly damped
it is very slow (on figure 2.9 axes are not orthonormal x− and y−wise), and its major component
on speed makes it easily controllable with a Vair feedback on thrust. On commercial airplane
phugoid damping is handled by auto-throttle or auto-thrust functions.

As a conclusion for this modes analysis, short-period instability is most problematic for han-
dling qualities. It was demonstrated that short-period approximation captures well the underlying
physics, hence most of next studies in this report will only focus on the [α, q] approximation,
making the assumption that the phugoid slow motion would be damped without problem by a
speed feedback on thrust.

2.4.2 Lateral Modes

Lateral dynamics is more similar to that of a conventional tail-aft airplane, still with some speci-
ficities. A main conclusion of preceding work [Saucez, 2013] was the necessity of adding vertical
surfaces to baseline BWB configuration in order to provide sufficient natural dutch roll damping.
A target of ξ = −5% natural dutch roll damping was obtained 7 when including two 10m2 vertical
surfaces, either placed over the engines or at wing tips. No proper OAD integration of these
lateral surfaces was performed since, so in our study two fins mounted over the engines were
considered.

2.4.2.1 Pure Roll Mode

Roll mode is associated with roll motion equation (2.45). It may be understood as a converging
mode of roll rate p when solicited by a step input of aileron – or elevon δmi in our case–.
Considering only the linearized roll motion equation gives:

ṗ = lp p + lδm δmi (2.67)

which corresponds to a first-order mode with time response τroll = − l1p , that is:

2mRxx
τroll = − (2.68)
ρVe 2 SlClp

For a given airplane and flight point, time response depends only on aerodynamic coefficient
Clp . This coefficient is a damping term corresponding to the opposing roll moment coming from
the lift difference between wings during a roll maneuver: if a right turn starts, the right wing sees
an increased local angle of attack compared to the left wing; as the result additional lift is created
7
assuming then a closed-loop stabilization
2.4. Open Loop Modes Analysis 59

Roll mode Dutch Roll


Spiral mode

Figure 2.10: Evolution of Lateral Poles with Speed.

on right wing that opposes the roll motion. Consequently Clp < 0 whatever the aircraft, and roll
mode is always stable.

Then equation (2.68) also shows that time constant diminishes when dynamic pressure in-
creases, which is corroborated by poles locations as a function of Mach for a given altitude
displayed on figure 2.10. Frequency varies with Mach from 0.7 to 1 rad.s−1 , which is rather slow
compared to a conventional airplane.

2.4.2.2 Spiral Mode

This mode is always real and slow: it corresponds to a convergence mode on bank angle φ . It may
be stable or not: the former case implying that, from an initial bank angle φi , the airplane tends
to get back to straight wings φ = 0˚, whereas in the latter case bank angle tends to increase. For
the considered airplane, the spiral mode is quasi neutral whatever the flight point, see figure 2.10,
which is a desirable behavior from a pilot perspective.

2.4.2.3 Dutch Roll Mode

This mode is an oscillatory motion on sideslip β and yaw rate r , with also an influence on roll
rate p through coupling term lβ of equation (2.55). However an usual approximation is to study
this mode characteristics only on reduced state-space [β, r]:
60 Chapter 2. Flight Dynamics Models

! ! ! ! !
β̇ yβ yr − cos α β y yδn ∆δmi
= + δmi (2.69)
ṙ nβ nr r nδmi nδn δn

Similarly to what was computed for short-period motion, following characteristic equation
results from state matrix:

s − yβ cos α − yr
|sI − Adr | =
−nβ s − nr
|sI − Adr | = s2 − s(yβ + nr ) + yβ nr − nβ (cos α − yr ) (2.70)

After identification with a second-order equation, following characteristics result:


2
ωdr = yβ nr − nβ (cos α − yr ) (2.71)
−(yβ + nr )
ξdr = (2.72)
2ωdr

Physical interpretation is not as straightforward as for longitudinal case. However it can be


observed that dutch roll damping is directly linked to Cyβ and Cnr aerodynamic coefficients,
which are both directly influenced by vertical surfaces area. That link indeed enabled in previous
work setting a target on dutch roll natural damping with lateral surfaces area as a degree of
freedom.

On figure 2.10 badly damped, and even negative dutch roll is observed, with a frequency around
0.4 rad.s−1 . Once again high instability comes with reduced dynamic pressure, that decreases
aerodynamic damping effects.

2.5 Turbulence Model

In order to evaluate the aircraft response under exogenous disturbance, a Dryden turbulence model
is used for temporal simulation. Since this study is mainly concerned with evaluation of longitudinal
instability, only vertical turbulence wz , expressed in R0 frame, is considered. Following formulas
from [Gage, 2003], Dryden turbulence model gives following power spectral density:
 
ω
σz 2 Lz 1 + 3 Lz Vair
ΦW z (ω) = 2 2 (2.73)
πVair
 
ω
1 + Lz Vair

Temporal simulation of this spectral density may be obtained by passing a white noise with
unit power ew through following filter, written in Laplace formalism:
2.5. Turbulence Model 61

Figure 2.11: σz as function of altitude and probability of exceedance, according to [Chalk,


1969] .


3Lz
s
wz 2Lz 1 + Vair s
HW z (s) = = σz 2 (2.74)
ew πVair 1 +

Lz
Vair s

Different values for turbulence length scale Lz and root-mean-square intensity σz exist in
literature, depending on altitude and probability of exceedance by flight hour. For instance US
military specifications [Chalk, 1969] provides values displayed on figure 2.11.

Concerning the choice of probability of exceedance, the rationale in our study is to focus on
sizing cases, meaning turbulence intensity jeopardizing the aircraft safety, rather than comfort
criteria. Therefore we only work with “severe” turbulence case, in order to verify that the aircraft
remains controllable even in those rare situations. Following literature review following values were
chosen:
62 Chapter 2. Flight Dynamics Models

Lz = 50m (2.75)
−1
σz = 5m.s (2.76)
Chapter 3

Review on Control Laws Techniques


and Application on an Academic
Example

Contents
3.1 Presentation of the Academic Example: the Inverted Pendulum . . . . 64
3.1.1 Open-Loop Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.1.2 Structure of the Stabilizing Control Law . . . . . . . . . . . . . . . . 67
3.1.3 Actuator Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.2 Eigenvalues Assignment through Analytical Computation . . . . . . . . 69
3.2.1 Analytical Computation Without Actuators Dynamics . . . . . . . . 69
3.2.2 Inclusion of Actuators Dynamics . . . . . . . . . . . . . . . . . . . . 70
3.3 Linear Quadratic Approach for Minimum Energy Control . . . . . . . . 74
3.4 H∞ Formulation with Weighting on the Acceleration Sensitivity Function 75
3.4.1 Introduction to H∞ Control . . . . . . . . . . . . . . . . . . . . . . 75
3.4.2 Acceleration Sensitivity Function (ASF) . . . . . . . . . . . . . . . . 77
3.4.3 Full-Order H∞ Compensator . . . . . . . . . . . . . . . . . . . . . 78
3.4.4 Structured H∞ Compensator . . . . . . . . . . . . . . . . . . . . . 81
3.5 Co-design Approach on an Inverted Pendulum . . . . . . . . . . . . . . 83
3.5.1 Optimal Actuator Bandwidth . . . . . . . . . . . . . . . . . . . . . . 83
3.5.2 Minimal Stick Size . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

As included in “Methods and Tools” part, this chapter is dedicated to describing control design
tools that were used throughout our study. First, specifications are listed, then several control
laws techniques are studied. Stabilization on an academical example illustrates each technique.
This academical example is also used to illustrate the “co-design” approach, i.e. synthesizing in a
single loop control laws and some meaningful physical parameters.

This thesis is by essence multi-disciplinary: a side intent of it is building bridges between


separate disciplines, namely flight mechanics, aircraft design and control. While the previous
chapter was devoted to describing flight mechanics models used in our work, this section lists
several control laws synthesis techniques for stabilizing an unstable plant. In order not to remain
only theoretical, those synthesis techniques are illustrated on a basic academic example: the
inverted pendulum. Integrated design and control 1 techniques are also applied on this simple
1
also sometimes referred to as co-design techniques. From now on both terms are employed interchangeably.

63
64 Chapter 3. Review on Control Laws Techniques

example. While such a simplified test-case may at first sight seem quite straightforward, we find
a deep pedagogical interest in this elementary approach. Indeed most phenomena at stake for
BWB design and stabilization are found on the inverted pendulum case, so we hope a careful
explanation on the academic example will ease understanding of latter parts on BWB control.

3.1 Presentation of the Academic Example: the Inverted Pendu-


lum

mp

θp

lp
y
+
Wp
x

up

Figure 3.1: Inverted Pendulum.

The inverted pendulum has been a research topic for control community since decades, see
for instance [Anderson, 1989]. It is indeed one of the simplest model of an instable plant one
can get, so it was used to model rockets, unstable airplanes [Stein, 2003], and with a variety
of variants such as 3D-pendulum for human walk modeling [Kajita et al., 2001], double inverted
pendulum [FURUTA et al., 1980], and inverted pendulum on a cart. In order to keep the example
as self-explanatory as possible, we restrict to the most basic model of torque-controlled inverted
pendulum, shown in figure 3.1. The choice of controlling the pendulum with a torque effector,
rather than an actuated cart for instance is not motivated only by simplicity considerations: it
is also a closer modeling of the aircraft dynamics, where pitching effectors such as elevators or
elevons create a pitching moment to control unstable pitch mode, as explained in section 2.1.3.
3.1. Presentation of the Academic Example: the Inverted Pendulum 65

3.1.1 Open-Loop Dynamics

Neglecting any friction force, and assuming that stick inertia is negligible compared to tip mass
inertia with respect to pivot point mp lp 2 , Newton’s second law reads:

mp lp 2 θ¨p = mp glp sin θp + up + Wp (3.1)

where θ¨p stands for the double time-derivation of angle θp with respect to an inertial frame. A
perturbing torque Wp is included in the model, accounting for any external torque that may move
the pendulum away from its unstable equilibrium. It may represent air friction, human disturbance
and so on.

Remark 3.1
In order to avoid inconsistencies between aircraft and pendulum parameters, all pendulum param-
eters are denoted with a subscript. Consequently:

• mp denotes the pendulum mass.

• θp denotes the angle between pendulum stick and vertical axis.

• lp denotes the stick length.

• up denotes the torque command.

• Wp denotes the perturbing torque.

Linearization of equation (3.1) around unstable equilibrium point θp = 0˚ and writing with
Laplace formalism gives: !
2 g 1
s − θp = (up + Wp ) (3.2)
lp mp lp 2

State-space representation of equation (3.2) with states X = (θp θ̇p )T gives:

! ! ! ! !
θ̇p 0 1 θp 0 0 up
= g + 1 1 (3.3)
θ̈p lp 0 θ̇p mp lp 2 mp lp 2 Wp

We assume that both pendulum angular position θp and speed θ̇p are measured. From
equation (3.2) two roots come directly:
s
g
s=± (3.4)
lp

q
The system features two poles, one stable and one unstable, both with frequency ωp = lgp .
It should be noted that this frequency does not depend on mass mp , and the shorter the stick,
the more unstable the pendulum. Following numerical values are used for demonstration purpose:
66 Chapter 3. Review on Control Laws Techniques

Pendulum mass mp 2 kg
Stick length lp q
1m
Eigenvalue frequency ωp = lgp 3.13 rad.s−1

Table 3.1: Pendulum numerical values.

Remark 3.2
There are strong analogies between equation (3.3) and equation (2.57), that is short-period
oscillation may be approximated as an unstable pendulum under following assumptions:

• Damping terms Czα and CmG q are neglected. Whereas CmG q origin and influence
was already discussed in section 2.4.1.1, Czα lift gradient is obviously not at all negligible
in itself. However not considering its influence on short period motion is equivalent to
considering that the airplane center of gravity remains fixed in the inertial system, or said
otherwise γ ≈ 0 during the motion. In this case α ≈ θ.

• Czq is neglected.

• Control surfaces act only on moment equation: Czδmi is neglected.

q
g
Following this analogy, we see that mα ⇐⇒ lp , and recalling that a necessary condition for
airplane static stability is a positive static margin, it immediately comes that the stick is unstable,
which comes as no surprise. Similarly, it was shown in section 2.4.1.1 that maximum longitudinal
instability was ω ≈ 1.3 rad.s−1 , which corresponds to a lp ≈ 6 m stick.

Pendulum Dynamics
g
lp

wp

θc − + utheo up + 1
+ θ̈p
1 θ̇p
1 θp
k1 Tact 2
mp lp
+ + + + s s

k2

Control Law

Figure 3.2: Block-diagram of pendulum dynamics and stabilizing control law.


3.1. Presentation of the Academic Example: the Inverted Pendulum 67

3.1.2 Structure of the Stabilizing Control Law

It can be shown easily, for instance through the root-locus diagram analysis of figure 3.3, that a
simple feedback of the angular position θp on the torque command up with a gain k1 is not
sufficient to stabilize the pendulum, and leads at best to an oscillatory motion. For this reason an
angular speed feedback with gain k2 is added to the control law. Now there are two independent
state-feedbacks with two gains, namely k1 and k2 , in order to place two modes. According
to linear control theory, these two degrees of freedom allow placing closed-loop pendulum modes
where desired. The control law is written as follows:

up = k1 (θp − θc ) + k2 θ˙p (3.5)

Root Locus
4

2 Increasing value of k1
-1)
Imaginary Axis (seconds

1 q
g
−ωp = − lp
0
q
g
ωp = lp
-1

-2

-3

-4
-4 -3 -2 -1 0 1 2 3 4
Real Axis (seconds-1)

Figure 3.3: Open-loop poles location of inverted pendulum and effect of a position feedback with
gain k1 .

Remark 3.3
In our example the control law only aims at stabilizing the pendulum, i.e. keeping θc = 0˚
despite any perturbations. For this reason θc may sometimes be omitted in further equations.
Furthermore additional goals such as reference input tracking would require a different control law
structure, for instance an error integral term.

Consequently we specify the desired closed-loop behavior under the form of following second-
order transfer function:
68 Chapter 3. Review on Control Laws Techniques

ω0 2
Wdes = (3.6)
s2 + 2ξ0 ω0 s + ω0 2

Said otherwise, we want the closed-loop system to behave as a second-order system with
frequency ω0 and damping ratio ξ0 . For this pendulum example, following values are chosen:

ω0 = 5 rad.s−1 (3.7)
ξ0 = 0.7 (3.8)

Classically, ξ0 = 0.7 offers good compromise between overshoot and fast time response, and
ω0 = 5 rad.s−1 aims at accelerating the closed-loop frequency compared to the open-loop.

3.1.3 Actuator Dynamics

−3dB

up ωact
utheo

1
ωact Time (s)
(a) Step response of a first order actuator with band- (b) Frequency response of a first order actuator with
width ωact = 5Hz. bandwidth ωact = 5Hz.

Figure 3.4: Temporal and frequency responses of a first-order actuator with bandwidth ωact =
5Hz.

The applied torque up physically originates from a mechanical actuator, e.g. an electrical
motor. The theoretical order utheo computed by the control law may not be in general directly
translated into the actual torque up : rather internal actuator dynamics delays settling time of the
actual order. This dynamics is taken into account through the transfer function Tact . In this
part actuator dynamics is either modeled as a perfect transfer Tact = 1, meaning control law order
utheo is immediately translated into the applied torque up , or as a first order transfer function
of bandwidth ωact :
up ωact
Tact = = (3.9)
utheo s + ωact
Temporal and frequency characteristics of such transfer function are displayed on figure 3.4.

More precisely, as pointed out by [Stein, 2003], every physical systems have a finite “available
bandwidth”, that is a frequency range beyond which their physical behavior is not well modeled,
3.2. Eigenvalues Assignment through Analytical Computation 69

understood or measured. For instance modeling our unstable pendulum required making implicit
assumptions: we assumed the stick is stiff, so flexible stick modes are neglected. We also assumed
that air friction was negligible. Sensors that measure values of θp and θ˙p also have a finite
frequency range, and so does the actuation loop converting measurements into actuators orders
utheo . Minimum of all these physical limitations may be referred to as “available bandwidth”, and
is precisely what Tact model is accounting for. However from now on we will refer to this available
bandwidth as actuator bandwidth ωact , assuming the actuator is limiting, yet acknowledging that
this terminology is restrictive compared to what it really models.

In order to assess the influence of actuator bandwidth Tact on different control laws perfor-
mance, two separate values for Tact are examined. Recalling that open-loop instability equals
ωp ≈ 3.13 rad.s−1 ≈ 0.5Hz, high and low bandwidth actuator are respectively defined with
respect to this value:

• Tact = 10 ωp = 5Hz. A common engineering rule indeed typically advises that the available
bandwidth should be ten times the dynamics to control in order to avoid any control issue.

• Tact = 3.5 ωp = 1.75Hz. A value of 3.5 is considered with the rationale that it should be
quite challenging for control, still without being catastrophic.

3.2 Eigenvalues Assignment through Analytical Computation

This simple example allows analytical computation of gains k1 and k2 so that exact location of
two poles may be achieved. Equivalent theory in general case – that is when analytical computation
is not tractable – is referred to as eigenvalues assignment by static output feedback, and is
developed comprehensively in [Bérard et al., 2012] and references therein.

3.2.1 Analytical Computation Without Actuators Dynamics

First no actuator dynamics is assumed. Open-loop dynamics equation (3.2) and feedback control
law of equation (3.5) put together result in following closed-loop equation:

θp −k1 /mp lp 2
= (3.10)
θc s2 − mkp2lp 2 − lgp + k1
mp lp 2

Denominator of equation (3.10) is then identified with the desired denominator from equa-
tion (3.6), giving following expressions for feedback gains:
!
2 g
k1 = −mp lp + ω0 2 (3.11)
lp
k2 = −2mp lp 2 ξ0 ω0 (3.12)
70 Chapter 3. Review on Control Laws Techniques

This gains computation guarantees that closed-loop poles will be at frequency ω0 and
damping ξ0 , at least provided pendulum dynamics is correctly modeled, and actuator dynamics
is neglected 2 .

Values computed with equations (3.11) and (3.12) are displayed on table 3.2. Obviously,
as no actuator dynamics is taken into account, gains values are the same whatever the actuator
bandwidth. Then responses to an initial impulse perturbation on Wp are displayed for ωact = 5Hz
and ωact = 1.75Hz on figures 3.5 and 3.7 respectively. It should be clear to the reader that
even though gains are computed without actuator dynamics taking into account, that is with
equations (3.11) and (3.12), actuators dynamics is included in the simulation: model displayed
on figure 3.2 is used for simulation. On figure 3.5 available bandwidth is much greater than the
dynamics to control, so closed-loop response with gains computed without taking this dynamics
into account stay relatively close to the ideal second-order response – which is superimposed
with full-order H∞ response, as it will be developed in section 3.4.3–. However when actuator
bandwidth is decreased, on figure 3.7, gains are no more satisfactory and a badly damped oscillatory
motion results. Similar results are visible on complex plane pole display of figures 3.6 and 3.8:
whereas for ωact = 5Hz poles remain relatively close to the ideal model ω0 = 5 rad.s−1 , ξ0 = 0.7,
for ωact = 1.75Hz closed-loop poles damping decreases a lot: ξ ≤ 0.4.

3.2.2 Inclusion of Actuators Dynamics

In order to improve the feedback quality, actuators dynamics is now taken into account for gains
k1 and k2 computation. For this simple example, analytical computation remains straightfor-
ward. Including a first-order actuator dynamics from equation (3.9) gives the following closed-loop
transfer function:
θp −k1 ωact /mp lp 2
= 3 (3.13)
θc s + ωact s2 − (g/lp + ωact k2 )s − (g/lp + k1 ωact )

Denominator of previous equation is then identified with desired closed-loop poles location,
given by:

(s + s1 )(s2 + 2ξ0 ω0 s + ω0 2 ) = s3 + (s1 + 2ξ0 ω0 )s2 + (2ξ0 ω0 s1 + ω0 2 )s + s1 ω0 2 (3.14)

where s1 is the third closed-loop pole location, coming from actuator first-order dynamics.
It is important to understand that with only two gains k1 and k2 , location of s1 cannot be
controlled: it is a result of design choice for pole placement of the two plant modes. Here we
choose to constraint two closed-loop modes to frequency ω0 and ξ0 exactly, and s1 is deduced
from this choice. As a consequence this technique gives no guarantee on closed-loop stability: for
a too low value of ωact , s1 may become unstable.

More precisely identification of equations (3.13) and (3.14) gives following values for s1 , k1
q
2 g
This is equivalent to supposing ωact  lp
.
3.2. Eigenvalues Assignment through Analytical Computation 71

and k2 :

s1 = ωact − 2ξ0 ω0 (3.15)


mp lp 2  
k2 = − g/lp + 2ξ0 ω0 s1 + ω0 2 (3.16)
ωact
 
k1 = −mp lp 2 g/lp + s1 ω0 2 /ωact (3.17)

From equation (3.15) we directly find that for values where ωact ≤ 2ξ0 ω0 , the third pole that
cannot be controlled turns unstable. Interestingly this value does only depend on the desired
closed-loop behavior and not on plant parameters 3 . For this example we find that the eigenvalue
assignment technique gives an unstable solution for ωact ≤ 1.1Hz.

Contrary to gains computation without actuator dynamics, significant differences between


gains values for ωact = 5Hz and ωact = 1.75Hz are observed in table 3.2. Poles displays in
figures 3.6 and 3.8 show an exact pole placement of two closed-loop poles on ω0 = 5 rad.s−1 , ξ0 =
0.7, which is what was aimed when computing gains. However the third closed-loop mode, so-
called s1 , moves significantly:

• ωact = 5Hz: s1 = 24.4 rad.s−1

• ωact = 1.75Hz: s1 = 4 rad.s−1

We remind that for a lower actuator value ωact = 1.1Hz this pole would turn unstable. On
temporal impulse response with “fast” actuator, figure 3.5 closed-loop response is close to the
optimal desired response. With “slow” actuator however response turns much slower, yet remains
stable and well damped.

ωact = 5Hz ωact = 1.75Hz


k1 (N m/rad) k2 (N m/ rad.s−1 ) k1 k2
Eigenvalue assignment, no actuator -174 -35 -174 -35
Eigenvalue assignment, actuator -146 -33 -95 -29
Linear-Quadratic -98 -31 -98 -31
Structured H∞ -151 -41 -125.4 -51

Table 3.2: Gains k1 and k2 computed with different control techniques, for two actuator
bandwidth.

Conclusion and Limitations As pointed out in [Syrmos et al., 1997], this compensation tech-
nique may result in an unstable plant, therefore the control designer has to be very careful in the
choice of the assigned dynamics and eigenvectors. A good knowledge of the plant to control is
required, otherwise wrong designs may result.
3
Yet this remark would not be true when considering a second-order actuator dynamics.
72 Chapter 3. Review on Control Laws Techniques

θp

θ˙p

Figure 3.5: Impulse Responses of controlled inverted pendulum with different control techniques
and ωact = 5Hz.

Figure 3.6: Poles of controlled inverted pendulum with different control techniques and ωact =
5Hz.
3.2. Eigenvalues Assignment through Analytical Computation 73

θp

θ˙p

Figure 3.7: Impulse Responses of controlled inverted pendulum with different control techniques
and ωact = 1.75Hz.

Figure 3.8: Poles of controlled inverted pendulum with different control techniques and ωact =
1.75Hz.
74 Chapter 3. Review on Control Laws Techniques

3.3 Linear Quadratic Approach for Minimum Energy Control

A widespread approach in control literature is the Linear Quadratic (LQ) compensator. Considering
initial conditions on states X(0), the objective is to find a stabilizing state-feedback up = KX to
bring back the states to zero while minimizing a cost index J:

Z ∞ 
J= X T QX + up T Rup dt (3.18)
0
This cost index stands for minimizing a weighted quadratic sum of state vector X and control
vector up . Q and R are weightings on states and control vectors respectively; their respective
size is n × n and m × m, n being the number of states and m the number of controls. Control
designer’s degrees of freedom move from gains selection to weighting matrices Q and R choice.
Trial-and-errors methods on these matrices coefficients enable making a compromise between
control performance and power consumption, for instance.

While the LQ approach is not restricted to linear problems, linear control theory proves the
uniqueness of a regulator minimizing the cost function J, and very efficient algorithms are known
to compute this regulator. Moreover stability and good margins properties are guaranteed with
this theory: the resulting regulator has at least 3dB gain margin and 60˚ phase margin. Kalman
introduced this theory in [Kalman, 1960]. For a comprehensive theory on optimal control, the
reader may refer to [Trélat, 2008].

However a well-known limitation of LQ regulators is that it is limited to state-feedback. To


overcome this limitation Linear Quadratic Gaussian (LQG) control theory was introduced so
that states are estimated using Kalman filter theory, but doing so robustness properties of LQ
regulator are lost. LQG theory is extensively described in [Alazard et al., 1999].

In our control benchmark a particular application of LQ control was used. Weighting matrices
are chosen as follows: Q = 0, R = 1. As a result performance index J simply becomes:

Z ∞
J= kup k2 dt (3.19)
0

This particular case is called minimum energy control. Indeed this choice of weighting
matrices leads to a performance index that only seeks minimizing control power required to stabilize
the pendulum. As in this case we have only one control, no degree of freedom is now left to the
designer. Interestingly it can be shown (see for instance [Alazard et al., 1999]) that the optimal
strategy to stabilize a plant with as few energy as possible is as follows:

• Stable modes remain at their initial location.


• Unstable modes are placed at their symmetrical stable counterpart with respect to the
imaginary axis.
q
g
For the pendulum case, we know from equation (3.4) that we have two poles located at ± lp .
3.4. H∞ Formulation with Weighting on the Acceleration Sensitivity Function 75

q
LQ theory guarantees that minimal energy control will result in two poles located at − lgp . This
minimal energy control is useful because all control techniques proposed in this chapter may have
their power consumption evaluated with respect to the optimum found by LQ regulator.

Gains are then computed for the pendulum example using Matlab lqr routine. Once again
state-feedback is assumed, and as we may not measure any actuator internal state, the gains are
computed on the plant without actuator while simulation incorporates actuator dynamics. For
ωact = 5Hz figure 3.6 shows one pole located at −3.13 rad.s−1 , and the other at −4 rad.s−1 ,
that is one pole is slightly moved by the introduction of actuator. Temporal response on figure 3.5
is well damped, but slow compared to desired second-order response. This is normal as no model
reference tracking was specified for LQ synthesis: slowness is the price to pay for minimum power
consumption. With a ωact = 1.75Hz actuator time response become slower and with greater
overshoot, but remains well damped. This is probably a feature of good robustness properties of
this technique.

3.4 H∞ Formulation with Weighting on the Acceleration Sensitiv-


ity Function

3.4.1 Introduction to H∞ Control

w z
P(s)

u y

K(s)

Figure 3.9: Standard form for H∞ control problem.

The H∞ control problem was formally introduced by [Zames, 1981]. Its most simple form is
defined through the standard form depicted on figure 3.9. Given a real rational transfer matrix
P (s), called plant, find a stabilizing controller K(s) such that an exogenous disturbance w acting
on the plant disturbs the predefined plant output z as few as possible, in the sense of a certain
metric defined hereafter. w , z , u and y are vectors in general Multi-Input Multi-Output
(MIMO) case. The H∞ control problem is basically a stabilization and perturbation rejection
problem. More precisely the general problem can be stated as follows:
76 Chapter 3. Review on Control Laws Techniques

min kTw→z (P, K)k∞ (3.20)


K∈K
such that: K internally stabilizes P

where Tw→z (P, K) is the closed-loop transfer matrix from external disturbance w to the
regulated output z , K is a controller space of rational transfer matrices K(s). k.k∞ is the
H∞ norm defined as follows:

kTw→z k∞ = max σ̄[Tw→z (jω)] (3.21)


ω∈R
kzk2
= max (3.22)
w6=0∈L2 kwk2

where σ̄(T ) represents the maximal singular value of T . For Single-Input Single-Output
(SISO) systems the H∞ norm corresponds to the peak response of the Bode magnitude diagram
over frequency. Minimizing the H∞ norm then is equivalent to minimizing the frequency peak
response for a SISO system. The latest equality is also of significant importance to understand
the rationale of H∞ control. The L2 norm defines the energy of a signal. For any disturbing
signal w with finite energy kwk2 , we seek to minimize the output energy kzk2 . More precisely
we seek the worst-case on the set of all finite-energy inputs.

Remark 3.4
The L2 -norm of a scalar function f : R → R is defined as:
Z ∞ 1
2
2
kf (t)k2 = |f (t)| dt (3.23)
0

For a MIMO transfer function F : Rm → Rp in frequency domain, the L2 -norm is defined as


follows: 
1 +∞  T
Z  1/2
kF (s)k2 = Tr G (−jω)G(jω) dω (3.24)
2π −∞
Perseval equality states the equivalence between the L2 -norms defined in temporal and frequency
domains.

Equation (3.20) then gives us a powerful and general formulation for a variety of control
problems. An adequate choice of disturbance w and output vector z allows shaping an
expected behavior on desired signal. For instance, if z is a tracking error, i.e. the difference
between reference and actual signals, then we want to keep this error as small as possible for
low frequencies. To do so we will weight the output z by a low-pass filter in order to minimize
kTw→z k∞ only for low frequencies. A different choice of z could also be used to minimize the
impact of high-frequency noise on command u . This general approach of shaping signals with
3.4. H∞ Formulation 77

weighting filters is called loopshaping, and is extensively described in [Zhou et al., 1996; Alazard
et al., 1999].

From an historical point of view, H∞ theory in control was considered as gathering the best of
two previous control approaches: from classical control theory it took frequency response analysis,
and more generally all control analysis tools from linear theory, and from so-called “modern” control
world it took both the powerful state-space formulation and optimal theory. The interested reader
may refer to [Aström and Kumar, 2014] for a comprehensive perspective on control theories.

3.4.2 Acceleration Sensitivity Function (ASF)

s2 +2ξ0 ω0 s+ω02
w s2 z

g
lp

+ + θ̈p θ̇p θp
up 1 1 1
Tact 2
mp lp
+ + s s

utheo y
P (s)

K(s)

Figure 3.10: Augmented Standard Form with Weighting on Acceleration Sensitivity Function
(ASF).

We now focus on a particular formulation of the H∞ problem: synthesis with weighting on the
Acceleration Sensitivity Function (ASF) . It was first introduced by [Fezans et al., 2007, 2008],
and later developed in detail in [Alazard, 2013]. The main interest of this approach is to construct
the H∞ problem directly from desired closed-loop specifications – in our case ω0 and ξ0 –
without any other tuning parameter such as plant parameters. In the case of MIMO systems,
dynamic decoupling between degrees of freedom is also easy to specify. This approach is well
adapted to problems formulated with second-order response, which is often the case of dynamic
systems actuated by force or moment, such as the inverted pendulum.

Augmented standard form with a weighting on the ASF is presented on figure 3.10. Dis-
turbance input w acts on the pendulum acceleration θ̈p , and the output z is the pendulum
acceleration θ̈p weighted by following weighting function, whose Bode magnitude diagram is visible
78 Chapter 3. Review on Control Laws Techniques

on figure 3.11:
s2 + 2ξ0 ω0 s + ω0 2
W1 = (3.25)
s2

Because of direct feedthrough value of 1 between w and z , we have:

kTw→z (P, K)k∞ ≥ 1, ∀K ∈ K (3.26)

More precisely optimal H∞ norm γ∞ with an optimal controller K∞ ensures that:

kTw→z (P, K∞ )k∞ = γ∞ ≥ 1 (3.27)


θp γ∞
⇒ (K∞ ) ≤ 2 , ∀s s.t. Re(s) > 0 (3.28)
w |s + 2ξ0 ω0 s + ω0 2 |

This simply means that a disturbance w will be rejected on pendulum position θp with a
dynamics at least equal to γ∞ times the desired dynamics s2 +2ξ0 ω10 s+ω0 2 , with γ∞ ≥ 1. Moreover
as the H∞ norm is a maximum, we know that for at least one frequency the closed-loop will match
γ∞ times the template.

3.4.3 Full-Order H∞ Compensator

The initial H∞ problem was solved using so-called “full-order” compensators. This denomination
is briefly explained hereafter. In fact solutions to the H∞ problem depend on the controller space K
where the optimal controller K∞ is sought. More precisely, optimal solutions to this problem were
initially sought in the space of controllers with a state-space form and with as many states as the
plant P (s), hence the denomination “full-order”. The rationale for this choice of solutions space is
that, as shown in [Gahinet and Apkarian, 1994], the problem becomes convex. As a consequence
different numerically efficient approaches were proposed: some were based on resolution of some
Ricatti equations, such as the DGKF algorithm introduced in [Doyle et al., 1989], others on LMI
solver, as described in [Gahinet and Apkarian, 1994].

For instance on our inverted pendulum problem, the pendulum features 2 states, and 1 state
is added with the actuators dynamics. A minimal realization of the augmented standard form of
figure 3.10 cancels the two integral poles introduced by weighting function P (s) 4 . As a result
P (s) is order 3, and the optimal compensator K∞ will also be order 3.

Full-order H∞ synthesis is then run on the inverted pendulum test-case, using hinfsyn routine
from Matlab, using Ricatti-based resolution method proposed in [Doyle et al., 1989]. As previously,
two actuator bandwidth ωact = 5Hz and ωact = 1.75Hz are compared.

For the “fast” actuator ωact = 5Hz, optimal H∞ value γ∞ = 1.001 is found, which means the
4
Proof of this assertation may be found in [Fezans et al., 2008], where a minimal realization of the ASF is
presented under block-diagram form. No pole is introduced by the weighting function.
3.4. H∞ Formulation 79

Peak response: 1.72dB at ω = 15.3 rad.s−1

Figure 3.11: Bode magnitude diagrams of inverse of frequency template 1/W1 , and ASF with
full-order and structured H∞ compensators, for ωact = 5Hz.

Peak response: 4.08dB at ω = 10.6 rad.s−1

Figure 3.12: Bode magnitude diagrams of inverse of frequency template 1/W1 , and ASF with
full-order and structured H∞ compensators, for ωact = 1.75Hz.
80 Chapter 3. Review on Control Laws Techniques

frequency template W1 is overshot by at most 0.1%. On figure 3.11 it is also visible that desired
sensitivity 1/W1 and optimal sensitivity kTw→z (P, K∞ )k∞ are superimposed, showing the ability
of ASF formulation to fit a desired closed-loop behavior.

Following compensator is found:

−7.7518.106 (s+1.508.105 )(s+31.42) −2.3511.1011 (s+31.42)(s+5.973)


h i
K∞ = (s+1.138.106 )(s+1.855.105 )(s+1) (s+1.138.106 )(s+1.855.105 )(s+1)
(3.29)

As expected, this compensator is order 3. We immediately see that it places a zero on the
actuator pole – that is on 5Hz ≈ 31.42 rad.s−1 . This is an undesirable behavior for it artificially
cancels the actuators dynamics by inverting it. While it may seem attractive, on a real system
where the available bandwidth is not exactly known, it would lead to poor robustness properties
of the law. Moreover this compensator features two very fast poles.

On figures 3.5 and 3.6 we see that full-order H∞ compensator exactly matches the desired
dynamics Wdes . Indeed among closed-loop poles, two are exactly placed on ω0 = 5 rad.s−1 , ξ0 =
0.7. Yet this is an artificially optimal compensator, because of compensating zeros and fast poles
previously mentioned. Reduction techniques could be applied to reduce the compensator order,
yet this is out of scope of this example. Additional filters could also be added to limit high-order
commands.

We get similar conclusions for the “slow” actuator case ωact = 1.75Hz: an optimal H∞ value
γ∞ = 1.001 is also found. Because of the actuator pole cancellation, the available bandwidth does
no more theoretically influence the closed-loop dynamics. Comparative Bode diagrams are shown
on figure 3.12. Once again the closed-loop ASF perfectly matches the template 1/W1 , which
results in a perfect closed-loop dynamics on figures 3.7 and 3.8. Following optimal compensator
results:

−7.7985.106 (s+1.508.105 )(s+11) −2.3653.1011 (s+11)(s+5.973)


h i
K∞ = (s+3.952.105 )(s+1.88.105 )(s+1) (s+3.952.105 )(s+1.88.105 )(s+1)
(3.30)

A third-order compensator is also found, and actuators dynamics ωact = 1.75Hz ' 11 rad.s−1
is also canceled with a zero.

Conclusion and Limitations While theoretically very powerful, the H∞ control approach re-
quires a certain amount of control engineering knowledge to produce implementable compensators
– see for instance [Puyou, 2005] for a multi-objective control law applied to an aeronautical appli-
cation –, getting rid of undesirable behavior such as high-order, high-bandwidth dynamics. And
obviously as soon as the optimal compensator is changed, the optimality certificate is lost.
3.4. H∞ Formulation 81

3.4.4 Structured H∞ Compensator

In order to retain the H∞ formalism while enabling the design of controllers more suited to
industrial applications, a research field emerged dedicated to optimizing structured controllers.
Structured means the control designer has the possibility to choose the controller order – which
is no more restricted to be the plant order – and which degrees of freedom inside the controller
the optimizer has access to. In our inverted pendulum application it simply means that we can
choose to optimize only position and speed feedback gains k1 and k2 in order to minimize the
H∞ norm kTw→z (P, K)k∞ .

Rationale for the advent of structured H∞ control is that in many industrial applications,
control designers already have a clear view on the system dynamics to control, and use very low-
order structure coming from their engineering knowledge, such as Proportional Integral Derivative
(PID) control, or C ∗ /Y ∗ structures for aeronautical applications, see for instance [Favre, 1994].
Instead of hand-tuning these gains, or applying classical control engineering tuning rules, structured
H∞ control now allows optimally computing these gains with respect to as many frequency criteria
as wished.

Yet if structured H∞ control has obvious practical advantages, adequate tools emerged only
recently. Optimization problem of equation (3.20) indeed becomes non-convex when controllers
space K is reduced to fixed-order, structured controllers. The subset of stabilizing controllers for a
given plant is typically non-convex, and may sometimes be disconnected, in parameters space K.
A second difficulty is nonsmoothness. The optimization objective of equation (3.20) is a min max
problem, informally meaning that we want to reduce as much as possible the peak response of the
transfer function. Yet a max function of a smooth 5 function is in general nonsmooth. What is
particularly problematic, is that the objective function nonsmoothness often appears at optimum
points of this function. Traditional gradient-based methods then cannot be applied directly to the
problem. Therefore new optimization tools had to be developed in order to handle nonconvex and
nonsmooth optimization.

Several concurrent methods were developed for H∞ optimization of structured controllers.


Among them is the hifoo package for Matlab, presented in [Burke et al., 2006]. Another
routine which is now included in Robust Control Toolbox from Matlab is hinfstruct, later
renamed systune after it can handle more diverse optimization problems. The optimization
methods used in these routines are explained in [Apkarian and Noll, 2006; Apkarian et al., 2007].
It uses Clarke subdifferential theory in order to compute local optima of equation (3.20). Local
approximations of the objective functions are then computed and descent directions are computed
for each iteration. More details on algorithms used as well as implementation peculiarities may be
found in[Gahinet and Apkarian, 2011b,a; Apkarian, 2012; Gahinet and Apkarian, 2013].

However it should be kept in mind that due to the nonconvex nature of the problem, these
algorithms converge only towards local optima: no guarantee of global optimality is given. To
overcome this difficulty, several initializations should be run in order to ensure that a correct
controller is found. Moreover no guarantee exists that for a given plant and a given controller
5
i.e. C ∞
82 Chapter 3. Review on Control Laws Techniques

structure there is a stabilizing solution. It is often argued that structured optimization aims at
integrating engineering knowledge into control optimization, so control designer is expected to
know in advance which controller structure works for a given plant dynamics.

In our example, the optimization problem becomes:

min kTw→z (P, k1 , k2 )k∞ (3.31)


(k1 ,k2 )∈R2

such that: [k1 , k2 ] internally stabilizes P

We use the same ASF form as presented on figure 3.10, with K(s) = [k1 k2 ]T , and
hinfstruct routine from Matlab. Once again actuator bandwidth ωact = 5Hz and ωact =
1.75Hz are compared. Gains values are given on table 3.2.

For ωact = 5Hz, the optimal H∞ value is γ∞ struct = 1.22, to be compared with 1.001 given by

full-order H∞ . We have a 22% error on frequency peak response compared to the “ideal” –yet
not feasible – optimal value of 1. Once again this is due to the fact that it is physically impossible
to perfectly match desired closed-loop response Wdes with a finite available bandwidth ωact and
a static controller. Please note that γ∞struct = 1.22 corresponds to the peak response of 1.72dB

of Bode magnitude response of figure 3.11, since 20 log(1.22) ' 1.72dB.

Closed-loop poles are presented on figure 3.6. The following closed-loop poles are obtained:

• 2 complex conjugate poles at ω = 5.48 rad.s−1 and damping ξ = 0.917.

• 1 stable pole at ω = 21.4 rad.s−1 .

Similarly for ωact = 1.75Hz, the optimal H∞ value is γ∞ struct = 1.63, which is consistent with

the 4.08dB peak response visible on figure 3.12. Obviously the optimal criterion is degraded, with
a slow available bandwidth desired model Wdes is still more out of reach.

Closed-loop poles, also visible on figure 3.8, are the following:

• 2 complex conjugate poles at ω = 9.17 rad.s−1 and damping ξ = 0.49.

• 1 stable pole at ω = 2 rad.s−1 .

Conclusion and Limitations Structured H∞ synthesis is an appealing approach, gathering


power of optimization tools with structured (static feedback in our case) controllers simplicity.
While in this example full potential of this method is not yet exploited – simpler tools provide
satisfactory solutions –, controller optimization provides a lot of possibilities that will be explained
more in detail in next chapter. Another theoretical advantage versus for instance pole placement
is that it guarantees finding a stabilizing controller –possibly after several initializations – if the
controller structure allows stabilization, which pole placement does not. However this may reveal a
weak argument for physical systems: here for actuator bandwidth below 1Hz structured synthesis
3.5. Co-design Approach on an Inverted Pendulum 83

would provide a stable system contrary to eigenvalue assignment, but resulting solution would
not be satisfactory, because too slow or oscillatory. There are physical limitations on system
performance that no method can avoid. Finally when using structured synthesis, the user should
keep in mind that output optima are only local: several initializations should be performed to
ensure a global solution.

3.5 Co-design Approach on an Inverted Pendulum

While previous part was dedicated to introducing several concurrent control law synthesis methods,
and in particular H∞ control, this part is devoted to applying previously described optimization
tools to two separate integrated design and control problems. More precisely, co-design of control
law and optimal actuator bandwidth is presented in section 3.5.1, and co-design of control law
and minimal stick size is developed in section 3.5.2. These two examples were chosen on purpose:
they both are simplified versions of integrated design and control studies carried on BWB on next
chapters. More precisely, chapter 5 will describe a method to size actuators bandwidth of a BWB
in an optimal way, taking into account energy consumption for stabilization, while chapter 7 will
explore optimal sizing of control surfaces with constraints on closed-loop amplitudes of controls
in maneuvers.

3.5.1 Optimal Actuator Bandwidth

Pa (s) s2 +2ξ0 ω0 s+ω02


w s2 z

g
lp

+ utheo + + + θ̈p θ̇p θp


k1 wact 1 up 1
2
1 1
mp lp
+ − s + + s s

k2

in out
wact wact wact

Figure 3.13: Block-diagram for co-design of control law gains k1 and k2 , and actuator bandwidth
ωact .
84 Chapter 3. Review on Control Laws Techniques

struct = 1.22
Let us state the following inverse problem: considering that optimal H∞ norm γ∞
found in section 3.4.4 for ωact = 5Hz is not acceptable compared to minimum theoretical value
γ∞ = 1, and knowing the influence of actuator bandwidth on this optimal value, which actuator
value gives 10% error on optimal H∞ norm, that is γ∞ struct = 1.1?

Of course a first approach could be to solve the problem iteratively: we increase the actuator
bandwidth as long as γ∞ struct > 1.1. For this simple problem, computational time would not be

an issue, but it would be different on a more realistic test-case. A second approach, called “co-
design”, consists in simultaneously optimize control law gains k1 , k2 and actuator bandwidth
ωact so that we find minimum bandwidth ωact – we seek to keep bandwidth as low as possible
– which entails no more than 10% error in optimal H∞ norm γ∞ struct .

Formally speaking, the problem simply reads:

min ωact (3.32)


(k1 ,k2 ,ωact )∈R3

such that: kTw→z (Pa (ωact ), k1 , k2 ) k∞ ≤ 1.1


[k1 , k2 ] internally stabilizes P

The main originality of this formulation is to combine a sizing problem with a stabilization and
control problem. This allows for closed-loop sizing of the unstable plant. Moreover equation (3.32)
is easily implementable using structured synthesis algorithms, more precisely hinfstruct’s evolu-
tion in Matlab, systune. systune implements the same algorithms as presented in section 3.4.4,
but with more diverse criteria specifications possibilities, such as H2 norms and LMI regions
constraints. Contrary to hinfstruct, which only allowed minimization problems, systune also
allows for minimization under constraints: norms to be minimized are called “soft” constraints,
and norms to be kept below 1 are called “hard” constraints.

Block-diagram for co-design problem is presented on figure 3.13. For sake of clarity tunable
blocks are depicted in orange, fixed parameters are in green, and weighting function is in blue. Such
conventions are consistent throughout this document. Block ωact is present twice, once alone:
this is a computational “trick” in order to specify minimization of ωact . Indeed systune only
supports norms minimization by default; yet obvious following relation holds: kTωin →ωout k∞ =
act act
ωact .

As a consequence equation (3.32) is translated into code as follows:

% soft constraint: minimization of w_act


Req_codesign=TuningGoal.Gain(’w_act_in’,’w_act_out’,1);
% hard constraint: h_inf norm below 1.1
ReqHinf=TuningGoal.Gain(’w’,’z’,1.1);
% syntax: systune(model, Soft, Hard)
systune(ST,Req_codesign,ReqHinf) ;
3.5. Co-design Approach on an Inverted Pendulum 85

Maximum gain as a function of frequency


10
Peak response: 0.833dB at ω = 21.8 rad.s−1
0

-10
Singular Values (dB)

-20

-30

-40

-50

-60

-70
10 -1 10 0 10 1 10 2 10 3
Frequency ( rad.s−1 )
opt 1.1
Figure 3.14: kTw→z (P, k1 , k2 , ωact )k∞ , in blue, and constraint W1 in yellow. Constraint is fulfilled
if blue curve is below yellow zone.

From this optimization procedure following values are obtained:

opt
ωact = 69.3 rad.s−1 (3.33)
' 11Hz
codesign
γ∞ = 1.1 (3.34)

Resulting sensitivity function is displayed on figure 3.14 together with weighting function
constraint W11 . Please note that once again, optimal H∞ value γ∞ codesign = 1.1 corresponds to

maximum peak response of 0.833 dB.


86 Chapter 3. Review on Control Laws Techniques

θp max = ±10˚

lpopt ?

Wpmax = 5N m

Figure 3.15: Illustration of co-design problem.

This simple example gives an overview of possibilities of co-design using optimization


for structured controllers: instead of providing a bound on maximum achievable per-
formance, the desired performance is specified and necessary condition for actuation
dynamics is deduced. Here we show that in order to remain within 10% error with
respect to theoretical H∞ norm achievable, we should increase the available dynamics
to 11 Hz.

3.5.2 Minimal Stick Size

A second example of co-design, perhaps more interesting from a designer point of view because
based on more physical considerations, is that of minimum stick size. It is a simplified form on
what will be developed in chapter 7.

Let us consider that we want to reduce the pendulum stick size as much as possible. There
could be several physical considerations for this: we may want to minimize mass, or space al-
3.5. Co-design Approach on an Inverted Pendulum 87

location. However there is a counterpart for reducing the stick size: the pendulum becomes
increasingly difficult to control. As a consequence of this, for a given disturbance torque Wp
, resulting pendulum angle θp will increase, even if regulated. The general problem is then:
how unstable can be a pendulum while remaining controllable and with not too much amplitude
on its displacements? We found that a good measure for this is to look at regulated output
level for calibrated disturbance. Here we see a strong analogy with the more general problem of
controlling an unstable aircraft: we may have benefits to go towards a more unstable aircraft as
seen in chapter 1, but at the expense of increased control magnitude and states levels for a given
disturbance such as turbulence. The question is then: how far can we go towards instability?

Pa (s) s2 +2ξ0 ω0 s+ω02


w s2 z

wp 1
lp g

+ utheo + up + + + θ̈p θ̇p θp


k1 wact 1 1 1 1 1 1
mp lp lp
+ − s + + + s s

k2

in out
lact lp lact

Figure 3.16: Block-diagram for co-design of control law gains k1 , k2 and minimum stick size
lp . Tunable blocks are colored in orange.

Remark 3.5
There is a significant limit to our analogy of minimum stick size for a pendulum and maximum
level of instability for an aircraft: if assuming that the pendulum is torque-controlled, such as
stated in equation (3.1), then the more stick size lp is reduced, the more torque control up
is effective (because torque up translates into angular acceleration through the inertia factor
1
mp lp 2
). More precisely, torque up will be acting at higher frequencies, but with lower magnitude,
when lp decreases. This is not at all the case for an aircraft: on the contrary if considering a
decrease of stability through moving the CG backward for a given aerodynamics, then control
surfaces lever arm diminishes and control level needs to be increased.

For this reason co-design approaches developed for pendulum and BWB are slightly different:
here we seek maintaining a maximum level of angle θp , whereas for BWB study case we will
seek maintaining a maximum level of control power u .

From this remark, we set two constraints to the co-design problem:


88 Chapter 3. Review on Control Laws Techniques

• We want to follow as closely as possible desired behavior Wdes , translated into an


H∞ constraint through the ASF form. We know that with gains feedback k1 and
struct . We want to keep regulation optimal, hence the constraint
k2 the optimal value is γ∞
kTw→z k∞ ≤ γ∞struct .

• For a disturbance torque Wp with value below 5N m, we want regulated angular position
θp not to exceed 10˚. As satisfying this constraint in temporal domain for any signal shape
of disturbance Wp is still an open research question, we restrict this constraint to be Root
Mean-Square (RMS) – that is on L2 norms. We want:

kθp k2 10˚
max ≤ (3.35)
Wp 6=0∈L2 kWp k2 5N m

From equation (3.21) we know that for SISO transfer from Wp to θp this is equivalent
to writing:
10˚
TWp →θp ≤ (3.36)
∞ 5N m

35
TWp →θp ∗ 5N m
kθp k2 (in degres) in response to kWp k2 = 5N m


30 kTw→z k∞

25

20
kTw→z k∞

15

10

1.22
0 opt
0.2 0.4 lp 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Stick length lp (m)

Figure 3.17: RMS value of angular position θp for disturbance Wp with RMS value of 5NM
struct after structured synthesis (in blue), for varying values of stick length l .
(in red), and γ∞ p

The minimum stick size codesign problem is then:


3.5. Co-design Approach on an Inverted Pendulum 89

min lp (3.37)
(k1 ,k2 ,lp )∈R3
struct
such that: kTw→z (Pa (lp ), k1 , k2 )k∞ ≤ γ∞
10˚
TWp →θp ≤ (= 2˚/N m)
∞ 5N m
[k1 , k2 ] internally stabilizes P

On figure 3.17 we provide a graphical representation of problem formulated in equation (3.37)


for a better comprehension. Structured H∞ synthesis, as proposed in section 3.4.4, is performed
struct is computed, as well as
for different values of stick length lp . Optimal sensitivity norm γ∞
H∞ norm of transfer between disturbance Wp and angular position θp . This value is plotted
multiplied by 5 N m. What is visible on the figure is then maximal RMS value of θp for
disturbance Wp with RMS value of 5 N m.

We see that optimal H∞ ASF norm γ∞ struct is constant for varying l . On the contrary, θ
p p
displacement strictly increases when stick size lp decreases. For large length values the curve is
rather flat, and when stick size becomes small – say below 1 m–, there is a dramatic increase of
angular displacement for equal disturbance. What we seek is optimal stick length lpopt for which
angular displacement is no more than 10˚. Constraints are represented as hatched zones. On the
figure we see that optimal value we seek is slightly above 50cm.

More precisely, equation (3.37) is implemented using systune routine as follows:

% soft constraint: minimization of lp


Req_codesign=TuningGoal.Gain(’lp_in’,’lp_out’,1);
% hard constraints: h_inf norm below gamma_struct
% and h_inf norm of transfer between wp and thetap below 10 \degres/5Nm
ReqHinf=TuningGoal.Gain(’w’,’z’,gamma_struct);
Req_deflection=TuningGoal.Gain(’wp’,’thetap’,5/rad2deg(10));
% syntax: systune(model, Soft, Hard)
systune(ST,Req_codesign,[ReqHinf,Req_deflection]) ;

Following optimization results are obtained, with ωact = 5Hz:

lp opt = 0.5294
k1 = −54.6N m/rad
k2 = −11.7N m/ rad.s−1
kTw→z (P, k1 , k2 , lp opt )k∞ = 1.22
TWp →θp (P, k1 , k2 , lp opt ) = 2˚/N m

90 Chapter 3. Review on Control Laws Techniques

Optimal stick value of 52cm is consistent with graphical explanation of figure 3.17.

Remark 3.6
Co-design procedure works well on this simplified example for one specific reason: for a given
actuator bandwidth ωact , the optimal H∞ norm of the ASF is independent from the physical
parameter to design, here lp . This is clearly visible on figure 3.17. Would it not be the case, then
from the two constraints only one would be limiting in general case: either constraint on ASF
norm is active, and closed-loop correctly matches desired dynamics, but deflection constraint is
not active and we could have reduced further lp without hitting the 10˚boundary. Or deflection
constraint is active, and closed-loop behavior is not necessarily satisfactory.

For general case when H∞ constraint becomes dependent from physical design parameter,
then workarounds need to be investigated. This will be developed in next chapters.

Remark 3.7
We should also remark that such co-design procedure involving simultaneous computation of gains
and physical parameters faces some alternatives from a designer perspective:

• If a relationship directly exists between control law gains and physical parameters, then
there is no need for a simultaneous optimization. In our example such relation exists, see
section 3.2.2. So we could have used directly these formulas, change stick length lp and
measure output θp for a given disturbance Wp . However such relations are only available
for simple example, and do not stand in general case.

• Another alternative would be to use loops or dichotomy procedures. Once again this could
make sense on this simplified example, yet would become increasingly costly as plant size
and complexity increases.

Conclusion

In this chapter several control laws techniques were investigated, and implemented on an academic
example: the inverted pendulum. H∞ synthesis, and more particularly H∞ optimization for
structured controllers, has been examined in detail. Using tools for nonsmooth optimization, two
integrated design and control problems are developed: the first problem consists in simultaneously
optimizing control law gains and actuator bandwidth in order to guarantee correct model tracking.
The second problem consists in optimizing stick length lp together with control law gains in
order to guarantee maximal pendulum angular deflection when facing a calibrated disturbance.
The so-called “co-design” approach shows very powerful for handling integrated design and control
problems, and will be continued in next chapters on more complex test-case of controlling a BWB.
Part III

SCIENTIFIC CONTRIBUTION

91
Chapter 4

Elaboration of Design Procedures for


Handling Qualities Sizing of an
Unstable Aircraft

This chapter aims at summarizing a handling qualities pre-sizing process for an unstable
aircraft. While several disciplines are known to be at stake for preliminary design of config-
urations such as BWB or more generally any unstable aircraft, whether longitudinally or
laterally, interactions and decoupling between disciplines were not yet properly formulated to
our knowledge. Our intent is then to address this global problem in a general way: what we
will describe should be applicable not only to BWB sizing, but also to any aircraft that
could suffer from natural instability: an aircraft with small VTP for instance, or a rotor-
craft. This process formalization is also a way to properly introduce the next chapters: first
an “iterative” process is described. It corresponds to the traditional way of working, involving
manual redesign and iterations when constraints are not met. “Codesign” procedures are
then presented as variants of this process: codesign of control surfaces size on the one hand,
and of actuators bandwidth on the other hand, are shown to be optimization improvements
over traditional disciplines sequencing.
Our main contributions in this chapter are:

• The formalization of a process for sizing an unstable aircraft from a handling qualities
perspective. Interactions of different disciplines, as well as their segregation on a time
scale, were not properly formulated to our knowledge.

• The derivation of this process by combining control and geometrical designs. This cou-
pled approach greatly reduces development time compared to the sequential approach.

Contents
4.1 Iterative Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.1.1 Aerodynamic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.1.2 Handling Qualities . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.1.3 Actuators Pre-Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.1.4 Systems Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4.2 Codesign Process for Control Surfaces Sizing . . . . . . . . . . . . . . 102
4.3 Codesign Process for Actuators Bandwidth Sizing . . . . . . . . . . . . 104

93
94 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

Handling qualities sizing for preliminary design is not adapted to sizing a naturally unstable
aircraft. Classical methods include volume coefficients, as described in [Roskam, 1995], trim and
equilibria in the whole flight envelope, and temporal maneuvers, possibly approximated by pseudo-
equilibrium. It is yet certain that modern airplanes are designed with smaller margins with respect
to instability than previous generations. For instance stability requirements for VTP sizing tend
to be relaxed: a target open-loop damping is set based on engineering judgment, implying a given
VTP size 1 . The underlying assumption is that in a more detailed design phase, control laws
engineers will manage to directionally control the aircraft with an appropriate command level: said
otherwise, the aircraft can be properly damped in the whole flight envelope without too much
rudder inputs. Huge knowledge on past conventional aircraft programs allows such an iterative
process, which up to now worked well in practice. Yet this process is not satisfactory from a
methodological point of view: if the fixed open-loop target instability level is found too ambitious
for control laws, it requires a costly redesign. On the contrary if it turns out too conservative,
better performance from the overall aircraft point of view may could been achieved. The only
satisfactory way of working would be to include controls directly into the design process.

A second example, similar to the VTP sizing case, and more linked to BWB problematic,
is that of longitudinal instability: to our knowledge, no proper backward criterion yet exists to
determine the limit for open-loop instability, even for a conventional aircraft. The only answer to
this question, developed for instance in [Calderara and Lemaignan, 2007], states that beyond a
certain point other backward handling qualities criteria become sizing, so there is no point seeking
for too much instability. This argument is valid for conventional aircraft, besides properly sized
ones: backward limits to the W&CG diagram should as much as possible be defined by more
than one criterion. This may not be true for unconventional configurations, especially for BWB
where instability may be very sizing. Besides, there is a methodological interest in defining a
closed-loop instability criterion, even though it may not be sizing.

Section 4.1 of this chapter is devoted to describing more in detail an iterative process for
handing qualities pre-sizing of an aircraft including control laws. By “pre-sizing” we mean that
this process is valid for future projects studies: at the end of the process it does not mean the
aircraft is ready to fly! Only the configuration is sane and de-risked, and deserves a detailed design.
Limits of this process are also highlighted. These limits lead to defining several alternatives to
the process in order to integrate previously separate tasks: the “co-design” process is therefore
explained as an enhancement of the traditional process, in sections 4.2 and 4.3.

4.1 Iterative Process

The initial process, designed as “iterative”, is depicted in figure 4.1. For sake of clarity, all
constraints and variables are not represented. Only most relevant constraints are mentioned,
to ease reading and comprehension. The process should be understood time-wise from top to
1
Of course concerning VTP sizing not only stability requirements are at stake: indeed VTP is often sized by
controllability criteria with one engine inoperative.
4.1. Iterative Process 95

Aerodynamic Model
Performance (CoP) Glider Control Surfaces
(Planform + Profiles) Area &
Instability
(Static Margin)
Definition Empennages

Handling Qualities
Open-Loop Actuators Dynamics
Max Deflection Handling Qualities: (Bandwidth)
(or CG range) Closed-Loop
Trim, Equilibria Control Laws Objectives
Handling Qualities: (Damping, Frequency,
Control Laws Decoupling...)
Perturbation
CG Range / Maneuver
( or Max Deflection)
Time

Max Deflection Max Deflection Rate

Actuators Pre-Sizing Hinge Moments


Max Deflection
Computation

Max Hinge Moment

Estimation of
Actuator Techno: Actuators Mass Actuator Techno:
EHA, EMA...
Actuators EHA, EMA...
Estimation
Power Consumption

Systems Sizing
Actuator Techno:
EHA, EMA... Failure
Actuator Techno: Cases
Risk / Probability

Systems Architecture
Control Surfaces Split

Figure 4.1: Iterative process. The time scale is increasing from top to bottom. Each box represents
a discipline of the process. Inputs for each discipline are represented by arrows coming to the
top of each box; outputs of each discipline are arrows coming from the bottom of each box.
Constraints for each discipline are coming to the side of each box.
96 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

bottom. Some of these constraints are expressed quite specifically for BWB case, but should be
adaptable to a more general case of unstable aircraft, or even rotorcraft, sizing.

4.1.1 Aerodynamic Model

First necessary step of this process is obviously to get a geometrical and then aerodynamic model
of the aircraft. Conceptually, this may be decomposed into two parts: on the one hand we need to
define a glider form – that is mainly planform and profiles –, and on the other hand we start with
baseline control surfaces position and area, and empennages. Concerning the BWB planform
definition , as discussed in section 1.1.2.2, various objectives and constraints need to be taken
into account for a proper sizing: maximum LoD , cruise deck angle, spanwise lift distribution...
here we restrict to two relevant constraints from stability and control point of view: position of
the CoP has a direct impact on aircraft trim drag, and instability level is defined by relative
position of CG and aerodynamic center XF .

Please note that we have said nothing about how to choose values for these parameters, as
well as control surfaces and empennages sizes, on purpose. These values are indeed inputs, or
constraints, of the problem, not results. This iterative process is abusively said “sizing”: it is
an evaluation process rather than a sizing process. For given inputs, such as instability level or
empennages sizes, we get evaluations of sizing values of interest, such as deflections amplitudes
in the whole flight domain, or actuators consumption. If those values exceed acceptable values,
then a re-design is necessary: instability needs to be decreased, control surfaces area increased
. . . and the process needs to be run again, until inputs entail acceptable outputs levels. In that
sense the process is said iterative.

4.1.2 Handling Qualities

4.1.2.1 Open-Loop Handling Qualities

Once an aerodynamic model is defined, handling qualities may be evaluated. We choose to


decompose this step into two separate phases. First, “open-loop” handling qualities criteria are
evaluated: it is basically checked that the aircraft is trimmable and with sufficient margin on
controls for the whole flight domain in terms of mass, Mach, configuration, and altitude. This
step gathers all criteria which do not require any feedback control laws, and often in preliminary
design phase consists mostly in equilibria resolution2 . The BWB open-loop handling qualities
for is typically investigated in [Saucez, 2013]. Two distinct ways of examining open-loop handling
qualities may be used:

• Either a CG range is fixed from W&CG studies, and necessary control amplitude is
2
Even dynamic maneuvers may be approximated as equilibria: for instance flight dynamics equations may be
solved in order to determine how much aileron deflection is needed to reach a target constant roll rate, without
consideration on dynamic phase to reach this prescribed roll rate.
4.1. Iterative Process 97

computed in order to be able to trim and control the aircraft in the whole flight envelope.
This mode is called “normal” mode.

• Or maximal amplitude of deflections is fixed, and resulting CG range for which the aircraft
is trimmable and with sufficient control authority is computed in the whole flight envelope.
This mode is called “inverse” mode.

Those two alternatives are represented schematically on figure 4.1. Such a procedure is not
restricted to unstable aircraft sizing: on the contrary, handling qualities sizing for conventional
and stable aircraft is mainly restricted to this phase.

4.1.2.2 Closed-Loop Handling Qualities

A second step, more specific to unstable aircraft sizing, arises later in a time scale: we call it
“closed-loop” handling qualities. It may be defined as examining control amplitude and maximum
rates from a set of temporal and dynamic maneuvers, with feedback control laws active. The
necessary inputs for this evaluation are basically:

• Hypothesis on actuators dynamics, for instance bandwidth. More generally a model of


acquisition chain, ie sensors, computers, filters, and delays, is required.

• Definition of control laws objectives: this is how we expect the aircraft to behave in closed-
loop. This includes targets in terms of damping, frequency, and decoupling between modes.

• We also need to specify under which external disturbance, or pilot maneuver, the dynamic
maneuver simulation is performed. This is fundamental, and this is the main point of
unstable aircraft sizing. As already seen for the inverted pendulum simplified case, if not
disturbed from its initial equilibrium position, a plant does not require any control, unstable
or not. Yet if it gets unstable, then for a prescribed disturbance, control amplitude and
control rate will increase – this is what we see on figure 3.17 –. So examining the control
level for counteracting a disturbance is a good measure of how much unstable we can be.

• A last input omitted on figure 4.1 is hypothesis on control law structure and technique.

Once again none of what is presented here is new. What is new is that we emphasize the
necessity of including this dynamic evaluation at preliminary design phase, which was not required
for conventional stable aircraft: such a task could be left for detailed design phase, knowing by
experience 1) that there would be no reason for control amplitude to be unexpectedly high and 2)
reaching saturation for control surfaces deflections or rates is not catastrophic on a stable aircraft.

Last point should be clear to the reader:

Hitting deflection stops or rate limits on an unstable aircraft is potentially catastrophic.


98 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

In such a case the aircraft reverts back to its open-loop, i.e. unstable, behavior. The reader
needing conviction on this may refer to [Stein, 2003]. As a consequence maximal deflections
and deflection rates resulting from predefined dynamic maneuvers should be compared to limit
values. If those values are exceeded, then a come-back to the “aerodynamic model” phase (see
figure 4.1), and a re-design of the configuration are necessary. It could mean decreasing maximal
instability, adding empennages, or increasing control surfaces size. Such loop is however costly,
and alternatives are developed in section 4.2.

Remark 4.1
As already mentioned, this process is aimed at being general enough to handle several instability
cases for different configurations. For instance if we seek to design an aircraft with reduced VTP
size and to compensate its lack of directional stability by yaw control with engines, then the
closed-loop handling qualities criterion will involve checking engines delta of thrust and thrust
derivative when facing lateral turbulence. Therefore “Max Deflection” and “Max Deflection Rate”
in figure 4.1 should more generally be understood as “Control Level” and “Control Derivative
Level”.

Remark 4.2
In this chapter only general philosophy of sizing criteria is given, in order to understand the
rationales made in our study. Detailed hypotheses on how these criteria were implemented for
BWB sizing cases, namely assumptions on actuators dynamics, turbulence and pilot maneuvers
models, and control laws structures and objectives, are detailed in dedicated chapters, namely
chapters 5 and 7.

4.1.3 Actuators Pre-Sizing

1 2
2 ρVair SCzδm δm

HM

Airframe

Actuator

δm

Figure 4.2: Schematic view of control surface actuator installation.


4.1. Iterative Process 99

Having checked that deflections amplitudes are compatible with target limits both in open-
and closed-loop in the entire flight envelope, a summarized conclusion at this stage is that the
aircraft basically can fly, at least with full control surfaces actuators operative, and with proper
information feeding normal laws. A second consideration is then: at which cost does it fly? As
already stated in section 1.1, a potential threat for BWB design, and more generally unstable
aircraft is indeed a risk of important actuator mass and power consumption, due to two combined
effects: large control surfaces leading to high hinge moments, and need for high deflection rates
to control instability.

4.1.3.1 Hinge Moments Computation

“Hinge moments” in the scope of this thesis inappropriately refers to aerodynamic moments relative
to control surfaces hinge, see figure 4.2. Their computation theoretically only requires knowledge
of the aerodynamic model. However this computation is often performed only during detailed
design phase of an aircraft project. Indeed sizing hinge moments are mostly located in “corners”
of the flight envelope, that is for high dynamic pressure cases. As a result many aerodynamic non-
linearities such as shocks or flow separations are typically at stake, strongly influencing maximal
hinge moments values. Yet at a future projects stage a reasonable hypothesis is to consider a
similar confidence in hinge moments evaluation than for the aerodynamic model used for handling
qualities evaluation.

Evaluation of sizing hinge moments is a necessary step for estimation of actuator power con-
sumption, as well as a common input for actuators mass models. Typical formulas for hinge
moments evaluation may be found in [Roskam, 1985c]. An estimation of maximum hinge mo-
ments for a HWB can also be found in [Garmendia et al., 2015a].

4.1.3.2 Estimation of Actuators Power Consumption

Instantaneous actuator power consumption is computed as follows:

˙
Pact (t) = δm(t) × HM (t) (4.1)

Actuator power consumption at time t, Pact (t), is equal to the product of control surface
˙
deflection rate δm(t) and instantaneous aerodynamic hinge moment HM (t) on this control sur-
face. However as already stated, dynamic aerodynamic moments are difficult to evaluate with few
information typical of future projects phase.

For this reason a common approximation for a first guess of actuator power consumption Pact
consists in introducing the notion of corner power PCP , based on knowledge on maximum hinge
moment HM max and maximum control surface deflection rate δm ˙ max .

More precisely a schematic view of typical actuator capability is presented on figure 4.3: feasible
˙ is depicted as a function of aerodynamic hinge moment on the control surface.
deflection rate δm
100 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

iso
-
Corner Power:

Ma
˙ max
PCP = HMmax ∗ δm

x
˙ max
δm PCP

Po
we
r

Actuator Power Consumption (kW)


Max Power
Deflection Rate ( rad.s−1 )

Actuator Capabilities

0
Hinge Moment (Nm) HM max

Figure 4.3: Schematic actuator capabilities curve: deflection rate δm versus hinge moment
HM .

Two characteristic points are visible:

˙ max the actuator can provide


• On the top left of the curve lies maximal deflection rate δm
without any load applied on the control surface.

• On the bottom right of the curve lies the maximum hinge moment HM max , or stall load,
the actuator can sustain statically, that is without being able to move the control surface.

A “maximum actuator capability curve” then joins these two points: from the stall load point,
if the load gets decreased, then the actuator is able to move the control surface more and more
quickly.

On this figure it results from equation (4.1) that iso-power curves are hyperbolas. The max-
imum actuator power is then obtained for a set (δm, ˙ HM ) located on the actuator capability
curve. A common way to approximate this point, found for instance in [Garmendia et al., 2014,
2015a], consists in introducing corner power PCP defined as:

˙ max
PCP = HM max × δm (4.2)

PCP is visible on figure 4.3: it represents theoretical power necessary to actuate the control
surface at maximum deflection rate δm ˙ max under stall load. This is a too demanding case for the
4.1. Iterative Process 101

actuator sizing: usually maximum deflection rate is required where dynamic pressure is low, that
is at low speed, whereas high hinge moments are met at high dynamic pressure.

Yet knowledge on existing actuators allows writing following rule:

Pact max ≈ 0.85 ∗ PCP (4.3)

To sum up from figure 4.1, necessary maximum deflection rate – estimated from stability
augmentation needs–, together with computation of aerodynamic hinge moments, provide a first
order of magnitude of the actuators consumption.

Remark 4.3
˙
Strictly speaking δm(t)×HM (t) is not the power provided by the actuator, but mechanical power
˙
resulting from control surface rotation with speed δm(t) under moment HM (t). This power is
indeed provided by one or several actuators, but an efficiency factor accounting for mechanical,
hydraulic and electrical energy losses within the actuator should be taken into account when
computing Pact . However considering other conservative assumptions such as notion of corner
power, hypothesis was made that these efficiencies may be omitted for preliminary evaluation.
From now on actuation power Pact is then considered equal to mechanical power of the moving
control surface. It also has the advantage of being independent from the number of parallel
actuators actually used to move the control surface.

4.1.3.3 Estimation of Actuators Mass

Once maximum hinge moment HM max and maximum required deflection rate are evaluated,
simple models for actuators mass may be used. Some, as in [Roos, 2003], do only take maximum
hinge moment as input to provide an actuator mass estimation. These mass models are based on
statistical considerations on existing actuators for aircraft control surfaces. More detailed mass
models for actuators may be found in [Garmendia et al., 2015a]. Actuators models are decomposed
into subcomponents, and each component mass is evaluated using statistical relations based on
existing actuator. These models require knowledge on maximum hinge moment needed, maximum
deflection rate, as well as geometrical considerations on control surface geometry.

Remark 4.4
Stricly speaking a correct estimation of the actuators mass would require knowing the number
of actuators per control surface. This choice is driven by safety considerations and depends on
control surface criticality (see for instance [Garmendia et al., 2015b], where according to elevon loss
criticality, number of actuators per elevon varies from two to four.) However in a first assessment
a linear relationship may be considered between actuator mass and power, as stated in [Wildschek,
2014]. As a consequence with this hypothesis there is a similar order of magnitude between a
single equivalent actuator and several parallel actuators. This hypothesis has the advantage of
enabling a first actuators mass estimation before a proper systems sizing.
102 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

4.1.4 Systems Sizing

Systems sizing from a flight control perspective involves defining a systems architecture in terms
of energy (hydraulic or electric) routing and segregation, power supply, actuators architecture and
technology ( Servo-Hydraulic Actuator (SHA) , EHA , EMA ...) to comply with safety analysis
in terms of risk and probability of failure cases. Major challenges of BWB configuration from
a systems design perspective are of course the high number of control surfaces, but also their
multicontrol functions: whereas usually a jammed aileron only affects roll capability, on the BWB
a jammed elevon implies a lack of trimming, pitch and roll control authority. Complexity of failure
cases tree is then further increased compared to classical aircraft architecture.

To our knowledge this problem is very little addressed in literature: in [Garmendia et al., 2015b],
capability of trimming the aircraft with each control surface jammed is evaluated. Following this
analysis, control surfaces which jamming is catastrophic –meaning it cannot be counteracted by
other control surfaces deflection – are equipped with four parallel actuators respectively. A second
study of interest is presented in [Belschner, 2011] and conclusions are summarized in [Kozek
and Schirrer, 2014]. A complete systems architecture for a BWB is proposed, based on the
assumption of EMA only. As with current EMA technology actuator jamming cannot be
guaranteed to be extremely improbable, the architecture involves one EMA per control surface,
and all elevons are splitted. The final design involves twelve redundant elevons on each side of the
aircraft. However to ease safety analysis, each elevon is either dedicated to roll or pitch control.
Yet chosen architecture is not entirely satisfactory: some failures lead to the aircraft loss with a
too high probability, in particular landing phase for some specific CG positions.

However, systems architecture and sizing was found to be out of scope of our work: we are
focusing on showing the aircraft is capable of flying from a handling qualities perspective, with a
predictably adequate actuators mass and power consumption.

4.2 Codesign Process for Control Surfaces Sizing

As already said, the iterative process described in section 4.1 has the major drawback of being
very costly if a redesign of control surfaces, empennages or glider form is necessary 3 . More
precisely, if closed-loop handling qualities evaluation shows unacceptable control level, the aircraft
geometric model needs to be changed, the aerodynamic model and flight control laws need to be
computed again. The last point in particular may turn out not straightforward: if control theory
recent advances aim at automatizing the process of control design, a straightforward method for
control laws computation from an aerodynamic model does not necessarily exists. Then we have
no insurance that proposed geometric changes in the aircraft configuration will match expected
control levels.

For these reasons we propose an improvement over the existing iterative process: the so-called
3
This is particularly true when no prior engineering experience exists for the aircraft, as it is the case for the
BWB .
4.2. Codesign Process for Control Surfaces Sizing 103

Aerodynamic Model
Performance (CoP) Glider Control Surfaces
(Planform + Profiles) Area &
Instability
(Static Margin)
Definition Empennages

Minimum Control Surfaces Size

Handling Qualities
Open-Loop Actuators Dynamics
Max Deflection Handling Qualities: (Bandwidth)
(or CG range) Closed-Loop Control Laws Objectives
Trim, Equilibria
Handling Qualities: (Damping, Frequency,
Decoupling...)
Control Laws Perturbation
CG Range / Maneuver
( or Max Deflection)
Time

Max Deflection and


Deflection Rate

Actuators Pre-Sizing Hinge Moments


Max Deflection
Computation

Max Hinge Moment

Estimation of
Actuator Techno: Actuators Mass Actuator Techno:
EHA, EMA...
Actuators EHA, EMA...
Estimation
Power Consumption

Systems Sizing
Actuator Techno:
EHA, EMA... Failure
Actuator Techno: Cases
Risk / Probability

Systems Architecture
Control Surfaces Split

Figure 4.4: Codesign process for control surfaces sizing. The main difference between this figure
and the iterative process description is that Maximum Deflections and Deflection Rates are now
constraints of the Closed-Loop Handling Qualities Process instead of outputs.
104 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

codesign process for control surfaces sizing, presented on figure 4.4. The main difference between
this process and the iterative process of figure 4.1 lies in the definition of the closed-loop handling
qualities criteria.

Now maximum deflections and deflection rates are no more outputs of the process,
but become constraints.

This means we specifically impose a target level for maximum deflections, typically driven by
geometrical considerations (usually ±30˚) and for maximum allowed deflection rates, imposed by
actuators technology.

Then we need to build a parameterized geometric and aerodynamic model, parameters being
the expected output of the process. In our case we seek the minimum control surfaces sizes that
fulfill closed-loop handling qualities assessment while matching control level targets. Building the
parameterized aerodynamic model enables a coupling between geometrical sizing, flight control
laws synthesis and closed-loop handling qualities evaluation. As a result in a single step we
guarantee an optimal control surfaces size that fulfills closed-loop handling qualities criteria.

This alternative process is philosophically identical to the codesign process of simul-


taneously optimizing the control law and pendulum length described in section 3.5.2.
It justifies a posteriori why we developed this academical example. Application of this
process to BWB elevons sizing is extensively described in chapters 6 and 7.

4.3 Codesign Process for Actuators Bandwidth Sizing

Another alternative to the iterative process of section 4.1 consists in simultaneously computing
flight control law and necessary actuator bandwidth in order to guarantee adequate closed-loop
stability and performance. This process is visible on figure 4.5. This time actuator characteristics,
namely its bandwidth, is no more an input for closed-loop handling qualities but an output of
the process. By simultaneously optimizing actuator bandwidth and stabilizing control laws we
guarantee an optimal sizing of the actuators for correct perturbation rejection, which is still an
open point of BWB design (see for instance conclusions of [Kozek and Schirrer, 2014]).

This process is fundamentally equivalent to that described in section 3.5.1 for the inverted
pendulum. Its application to BWB problem is exposed in chapter 5.

Conclusion

In this chapter different disciplines and their interactions were presented concerning the problem of
unstable aircraft sizing at preliminary design phase. Limits of an iterative process were highlighted,
and two alternatives were proposed. First a simultaneous design of geometric parameters and
flight control laws with constraints on control level is presented. A second alternative consists in
4.3. Codesign Process for Actuators Bandwidth Sizing 105

Aerodynamic Model
Performance (CoP) Glider Control Surfaces
CG (Planform + Profiles) Area &
Instability
(Static Margin)
Definition Empennages

Handling Qualities
Open-Loop Control Laws Objectives
Max Deflection Handling Qualities: (Damping, Frequency,
(or CG range) Closed-Loop Decoupling...)
Trim, Equilibria
Handling Qualities: Perturbation
Control Laws / Maneuver

CG Range
( or Max Deflection)
Time

Actuators Dynamics Max Deflection and


(Bandwidth) Deflection Rate

Actuators Pre-Sizing Hinge Moments


Max Deflection
Computation

Max Hinge Moment

Estimation of
Actuator Techno: Actuators Mass Actuator Techno:
EHA, EMA...
Actuators EHA, EMA...
Estimation
Power Consumption

Systems Sizing
Actuator Techno:
EHA, EMA... Failure
Actuator Techno: Cases
Risk / Probability

Systems Architecture
Control Surfaces Split

Figure 4.5: Codesign process for actuators bandwidth sizing.


106 Chapter 4. Handling Qualities Sizing of an Unstable Aircraft

simultaneously designing flight control laws and optimizing actuator bandwidth in order to provide
adequate stabilization. While this chapter is deliberately general, aiming at giving guidelines for
unstable aircraft preliminary sizing, next chapters will be devoted to application of the two proposed
alternative processes to BWB closed-loop sizing.
Chapter 5

Combined Optimization of Stabilizing


Longitudinal Control Law and
Actuators Bandwidth through H2/H∞
Synthesis

This chapter aims at conjointly optimizing actuators of trailing edge control surfaces together
with a longitudinal stabilizing control law, using a H∞ synthesis model with weighting on
the Acceleration Sensitivity Function (ASF). The optimization criterion is not directly on the
actuators bandwidth, but is on the mechanical power consumption for stabilizing the aircraft,
expressed as a H2 norm. First an analytical study of relative influence of control law gains and
bandwidth on actuators power consumption is performed, through their influence on deflection
rate. Then numerical codesign process is setup: for that purpose a synthesis model based on
the ASF scheme is developed. A H2 /H∞ design is proposed, in order to minimize power
consumption under acceleration disturbance rejection constraint. Computation of maximum
hinge moments per control surface, necessary for mechanical power computation, is briefly
described. Finally actuators bandwidth is parameterized, and optimized on the whole flight
domain.
Our main contributions in this chapter are:

• Adapting the ASF scheme for the longitudinal flight control design.

• Recovering numerically the analytical solution of the control allocation problem, pro-
vided by the pseudo-inverse solution.

• Formalizing the codesign problem as a H2 /H∞ optimization problem.

• Developing an optimization criterion which is directly linked to the actuators power


consumption.

• Sizing the actuators bandwidth on the whole flight envelope.

Contents
5.1 Analytical Influence of Control Laws Gains and Actuators Bandwidth
on Deflection Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

107
108 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

5.1.1 State-Space Representation of Different Models . . . . . . . . . . . . 109


5.1.2 Full Linear Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5.1.3 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.2 Synthesis of Stabilizing Control Law based on the Acceleration Sensi-
tivity Function (ASF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.2.1 Description of Synthesis Scheme with Weighting on the Acceleration
Sensitivity Function . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.2.2 Control Law Gains Structure . . . . . . . . . . . . . . . . . . . . . 117
5.3 H∞ Synthesis with Optimization of Allocation Gains . . . . . . . . . . 118
5.4 H2 /H∞ Synthesis with Optimization of Allocation Gains . . . . . . . 121
5.5 Codesign of Control Law Gains and Actuators Bandwidth with Cost
Function on Deflection Rate . . . . . . . . . . . . . . . . . . . . . . . . 126
5.6 Evaluation of Maximum Aerodynamic Hinge Moments . . . . . . . . . 130
5.7 Codesign of Control Law Gains and Actuators Bandwidth with Cost
Function on Actuators Power Consumption . . . . . . . . . . . . . . . 130
5.8 Actuators Bandwidth Sizing on the Whole Flight Envelope . . . . . . . 135

This chapter is dedicated to presenting a method for optimizing actuators power consumption
on a BWB design while ensuring closed-loop longitudinal stabilization. It is a direct application
on BWB design of the codesign process described in section 4.3, and with a similar methodology
to that presented in section 3.5.1 for the inverted pendulum example. The reader may refer to
these parts of the report for a comprehensive explanation of the rationale of this optimization
using control theory tools. Let us just remind that this chapter aims at providing answers on two
open points of BWB design, that are linked one to another. First, which dynamics should the
actuators have, provided they need to move unusually large control surfaces, at a possibly high
rate for stabilization purpose? Secondly, what would be the cost of such active stabilization in
terms of flight control systems actuators, knowing that actuator power consumption is directly
linked to control surfaces deflection rate as shown in section 4.1.3.2?

To provide answers to these open points, we will first analyze the analytical influence of flight
control laws gains and actuators bandwidth on deflection rate, using linear control theory tools.
Then a control synthesis scheme based on a weighting of the ASF will be developed, similarly to
what was proposed in section 3.4.2. Two processes are compared: minimization of an H∞ criterion
on the one hand, and minimization of a H2 criterion with constraint on a H∞ norm on the other
hand. For these two processes not only flight control laws gains are computed, but also allocation
gains, that is how command is “spread” among all control surfaces.

Next step consists in simultaneously optimizing flight control laws gains together with actuator
bandwidth, in order to minimize the influence of an exogenous disturbance on control surfaces
deflection rates while ensuring proper stabilization. A similar process is finally presented, that aims
at minimizing the influence of exogenous disturbance on actuator power consumption – instead of
seeking minimization of deflection rates. Such criterion requires a priori computation of maximum
hinge moments. How these values are computed will also be addressed in this chapter.
5.1. Analytical Influence of Control Laws Gains and Actuators Bandwidth on Deflection
Rate 109
5.1 Analytical Influence of Control Laws Gains and Actuators Band-
width on Deflection Rate

In this section the influence of both flight control laws gains and actuators bandwidth is studied.
For that purpose linear analysis tools are used: namely state-space representations of the aircraft,
actuators, control law and turbulence are written in section 5.1.1. Then a closed-loop state-space
representation is obtained by gathering all models. Perturbation of the aircraft states is considered
either with turbulence input, or initial values for states.

5.1.1 State-Space Representation of Different Models

5.1.1.1 Aircraft Model

Superscript a is used to write aircraft state-space representation. General state-space form reads:
(
Ẋa = Aa Xa + Ba Ua + Baw Uw (5.1)
Ya = Ca Xa + Da Ua + Daw Uw (5.2)

In this chapter only longitudinal stabilization is addressed, thus state vector Xa and output
vector Ya are the ones introduced in equation (2.39):

State-space matrices Aa , Ba , Ca and Da respectively are provided from equation (2.40),


with the minor difference that in this chapter we are concerned only with longitudinal stabilization
around an initial equilibrium. As a consequence only stabilization with elevons is considered, for
a constant engines thrust. Then control deflections vector Ua only reads Ua = (∆δmi )Ti=1..10 .

Concerning wind influence on aircraft states Xa , only vertical wind wz is considered, so we


have Uw = wz . Moreover vertical turbulence is approximated as an increase of angle of attack
due to the wind δαw . This angle of attack is computed as follows:
wz
δαw = arctan (5.3)
Ve

Linearized equation (2.40) adding an angle of attack increase due to turbulence δαw gives:

δ α̇ = −zV δVair − zα (δα + δαw ) + (1 − zq )q (5.4)

Linearizing equation (5.3) and combining with equation (5.4) finally gives:
 
0
−z /V 
Baw =  α e (5.5)
 
 0 
0
110 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

Similarly combining equation (2.41) together with equation (5.3) gives:


 
0
 . 
Daw =  .. 

 (5.6)
zα /g

5.1.1.2 Actuator Model

General state-space representation for actuators model is written as follows, with a u subscript:
(
Ẋu = Au Xu + Bu u (5.7)
Ua = Cu Xu + Du u (5.8)

Please note the difference between u , the control vector coming from the control law, and
Ua the actual control deflection vector, which is an input of the aircraft model. u could also be
viewed as a “theoretical” order, while Ua is the “actual” control deflection, difference between
u and Ua coming from a potential actuators internal dynamics.

• If no actuators dynamics is considered, then Au = Bu = Cu = 0, Du = I, and Ua = u.

• If a first order actuator dynamics is taken into account as in equation (3.9) for each control
surface, then:  
−ωact 1 . . . 0
 . .. .. 
Au =   .
. . . 
 (5.9)
0 . . . −ωact 10

 
ωact 1 . . . 0
 . . .. 
Bu =  .
 . . . .   (5.10)
0 . . . ωact 10
 
1 ... 0
. . .. 
Cu = 

.
. . . . (5.11)
0 ... 1

Du = 0 (5.12)

Remark 5.1
In this study, separate control surfaces are allowed to have different actuator bandwidth. Never-
theless symmetrical control surfaces should have the same actuators. For this reason from now
on we set ωact i+5 = ωact i , ∀i = 1 . . . 5.
5.1. Analytical Influence of Control Laws Gains and Actuators Bandwidth on Deflection
Rate 111
5.1.1.3 Control Law Model

In this chapter we focus on the synthesis of a stabilizing longitudinal control law only: said oth-
erwise the problem is a disturbance rejection problem. Command tracking and lateral-directional
flight control laws synthesis will be addressed in chapter 7. More precisely we are working with a
static output-feedback compensator, whose general form is:

u = KYa (5.13)

It is a well-known fact that stabilizing and damping the short-period mode of an aircraft
basically requires a feedback proportional to α for stabilization and a feedback proportional to
q for damping – see for instance previously developed analogy with the inverted pendulum in
chapter 3, and control laws description in [Favre, 1994]. Yet in civil aircraft flight control systems
feedback of load factor nz is often preferred to α , for several reasons. The first reason is the lack
of confidence on anemometric measurements compared to inertial measurements: α measure
relies on rotating probes that are prone to icing, breaking or calibration errors, whereas inertial
measures are based on accelerometers within the aircraft. Then it is easier to know equilibrium
load factor in cruise, which is basically equal to one, whereas equilibrium angle of attack αe needs
a computation based on knowledge of aerodynamic and inertial parameters. But maybe the main
reason is historical: in FbW aircraft pilot stick input is a load factor input: pilot orders directly
affect the trajectory of the aircraft. As a consequence nz feedback may have seemed natural.

Consequently two feedbacks of nz and q , with associated gains Knz and Kq respectively,
are considered in this study. Moreover a third feedback in pitch attitude θ with a gain Kθ
needs to be added in order to be able to set the problem on the ASF form. More precisely in
order to apply the H∞ synthesis model with weighting on the ASF as developed in section 3.4.2,
we need to put the aircraft dynamics under a second-order differential equation form, which the
aircraft is not exactly. As a workaround we need to work only with three states [δα q θ], implying
an integrator pole coming from kinematic equation θ̇ = q. In order to be able to stabilize this
third pole a third feedback in θ is then required. This point, which is a drawback of the ASF
for our problem, will be discussed more in detail in section 5.2.2.

Finally all three feedbacks are summed up, and resulting order is spread among all control
surfaces through an allocation gain Kalloc . More precisely symmetric control surfaces are set to
move symmetrically:

h iT
1 5 1 5
Kalloc = Km . . . Km Km . . . Km (5.14)

Finally static output feedback is written as follows:


112 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

 
Km1
 . 
 . 
 . 
 5 h
Km  i
u=
K 1  0 0
 Kq Kθ Knz Ya (5.15)
 m
 . 
 . 
 . 
5
Km

5.1.1.4 Wind Model

General state-space representation of turbulence model is denoted with subscript w as follows:

(
X˙w = Aw Xw + Bw ew (5.16)
Uw = Cw Xw (5.17)

More precisely equation (2.74) is converted into state-space representation as follows:

" #
0 1
Aw = (5.18)
−Ve /Lz −2

" #
0
Bw = (5.19)
1

h q q i
Cw = σz Ve 3Lz (5.20)
πLz σz πVe

5.1.2 Full Linear Model

After having described each model separately, the models are now gathered in one unique state-
space representation.

First we gather the open-loop aircraft with the actuators model. Combining equations (5.1)
and (5.7) with hypothesis of first-order actuators gives the following state-space representation:

 ! ! ! ! !
 Ẋu A u 0 X u B u 0
= + u+ wz (5.21)



 Ẋa
 Ba Cu Aa Xa 0 Baw
!
   X
 u

 Ya = Da Cu Ca + Daw wz (5.22)
Xa


5.1. Analytical Influence of Control Laws Gains and Actuators Bandwidth on Deflection
Rate 113
Then adding control laws from equation (5.15) gives:

 ! ! ! !
 Ẋu Au + Bu KDa Cu Bu KCa Xu Bu KDaw
= + wz (5.23)



 Ẋa
 B a Cu Aa Xa Baw
! ! ! !
 Ua
 Cu 0 Xu 0

 = + wz (5.24)
Ya D a Cu Ca Xa Daw

Finally adding the wind state-space representation gives:


 X˙w
     

 Aw 0 0 Xw Bw
 Ẋu  = Bu KDaw Cw Au + Bu KDa Cu Bu KCa   Xu  +  0  ew (5.25)

       


 Ẋa

Baw Cw Ba Cu Aa Xa 0
    
 wz Cw 0 0 Xw



  Ua  =  0 Cu 0   Xu  (5.26)

     


 Y D C D C C X
a aw w a u a a

5.1.3 Analysis

For physical considerations it is easier to reason with vertical wind wz as an input of the closed-
loop aircraft, rather than white noise input ew . We then consider equation (5.23) instead of
equation (5.25). We also add a last output of interest: control surfaces deflection rates, which
we know to be directly proportional to power consumption for stabilizing the aircraft. We simply
add an output Uad = Ẋu . The closed-loop aircraft state-space representation becomes:

 ! ! ! !
 Ẋu Au + Bu KDa Cu Bu KCa Xu Bu KDaw
= + wz (5.27)






 Ẋa B a Cu Aa Xa Baw
     
Ua Cu 0 ! 0

 Xu
U = A + B KD C B KC + B KD aw  wz (5.28)
      
 ad u u a u u a u
X
     
a

 Y

Da Cu Ca Daw
a

Let us have a look at the initial response in terms of control surfaces deflection rates to a
vertical wind disturbance wz0 . At t = 0 all states are zero so we simply have:
114 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

Uad (t = 0+ ) = Bu KDaw wz0 (5.29)


 
ωact 1 Km
1
 .. 

 . 

ωact 5 Km5

 zα 0
ω 1 K 1  Knz g wz
= (5.30)

 act m 
 .. 

 . 

ωact 5 Km
5

This means i-th elevon deflection rate δm ˙ i at t = 0 is only influenced by i-th actuator band-
i i
width ωact , longitudinal allocation gain Km and control law feedback gain Knz . Here only nz
feedback has an influence because vertical turbulence only has a direct influence – said otherwise
a direct feedthrough – on the aircraft vertical load factor.

Similar conclusion can be obtained when examining initial deflection rate in response to an
initial condition on the aircraft states Xa0 . This initial condition is basically a way to simulate an
exogenous disturbance, for instance an increase in angle of attack. Everything else being equal to
zero we get:

Uad (t = 0+ ) = Bu KCa Xa0 (5.31)


 
ωact 1 Km
1
 .. 

 . 

 5 5 
ωact Km  h i
=
ω 1 K 1  Kq
 Kθ Knz Ya0 (5.32)
 act m 
 .. 

 . 

ωact 5 Km
5

Once again we see that for this particular case, the initial deflection rate is proportional both to
actuators bandwidth and to control laws gains. However one should not misunderstand what was
just demonstrated: equivalence between actuator bandwidth and control laws gains only holds for
initial response to perturbation. This partial equivalence comes from the fact that in state-space
representation 5.27 actuators bandwidth located in Bu matrix and control law gains located in
K are always present as a product Bu K and not separately. However ωact is also present
through matrix Au in the integration of actuators states. As a consequence tuning bandwidth
and control laws gains have dissimilar effects in general.
5.2. Synthesis of Stabilizing Control Law based on the Acceleration Sensitivity
Function (ASF) 115
It results that finding a compromise between adequate control laws tuning for proper
stabilization, and actuators bandwidth pre-sizing ensuring both good disturbance rejec-
tion and minimal power consumption seems out of reach with an analytical approach.
For this reason we propose to use numerical tools in order to simultaneously optimize
actuators bandwidth and control laws gains with constraints on closed-loop disturbance
rejection and minimum power consumption.

5.2 Synthesis of Stabilizing Control Law based on the Acceleration


Sensitivity Function (ASF)

In this part a numerical method is setup in order to simultaneously optimize control laws gains
and actuators bandwidth, with a cost index on actuators power consumption. This is developed
in several steps: first the H∞ -based optimization with weighting on the ASF , presented in
chapter 3 is adapted to the problem of longitudinally stabilizing the BWB . Necessary assumptions
for applying the ASF form on the longitudinal model are discussed, in particular constraints it
imposes on gains structure. Then a first synthesis based only on H∞ criterion minimization is run
and analyzed in section 5.3. An improved problem formulation is then proposed in section 5.4, that
avoids drawbacks of pure H∞ -norm minimization. Finally two codesign procedures are presented:
the first one aiming at minimizing the impact of an exogenous torque disturbance on actuator
rate, the second one aiming at minimizing the same disturbance on actuators power consumption.
These codesign procedures are extensively described in sections sections 5.5 and 5.7 respectively.

5.2.1 Description of Synthesis Scheme with Weighting on the Acceleration Sen-


sitivity Function

H∞ synthesis using the ASF form was introduced in section 3.4.2. We remind that this for-
mulation is well-adapted to controlling mechanical systems that can be modeled with generalized
second-order differential equations, such as spring-mass systems, or the inverted pendulum. For
this class of systems the ASF provides a simple and powerful way to specify desired closed-loop
behavior in terms of frequency and damping, for rejection of exogenous torque disturbance.

Motivated by the expected easiness of the ASF method, it was seeked to apply it on the
problem of longitudinal disturbance rejection for the BWB . However longitudinal flight dynamics
are not exactly a spring-mass system, so the analogy revealed limited. The only way to put
longitudinal equations into the form of a second-order differential equation is to consider θ as
the main variable. Therefore we get:


 θ̇ = q (5.33)


q̇ = mα δα + mq q + mδmi δmi (5.34)


α̇ = zα δα + (1 − zq )q (5.35)

116 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

q √
3Lz
ew 2Lz 1+ Ve s wz −zα
σz πVe (1+ Lz s)2 Ve
Ve

Turbulence Model
w s2 +2ξω0 s+ω02
zinf
s2

δαw
+ α
δ α̇ + δα
q
+
Ua Ẋa q̇ 1 q Xa Ya θ
Ba + +
s Ca +
nz
+ δ θ̇ δθ +

Aa
Da
Longitudinal Aircraft Model

+ Kq Delay
1 Uad i u
s ωact +
Kalloc δmequi
+ Kθ Delay

+ Knz Delay

Actuators Control Law


P (s)

Figure 5.1: Block-diagram of stabilizing control law computed with ASF scheme. Fixed param-
eters are in green, and tunable parameters are in red.

The α−equation is necessary in order to retain short-period dynamics. Yet adding θ to classical
2 × 2 formulation of short-period – considering only states [δα q] has a drawback: it adds a third
integrator pole to the two short-period poles.

Remark 5.2
In this chapter we only focus on the problem of exogenous disturbance rejection, that is the
problem of counteracting an unknown perturbation, as opposed to the problem of command
tracking, that is following a known input. The underlying assumption is that, especially for an
unstable configuration, it is harder –therefore more sizing – to reject something which is unknown
than something known. Yet the problem of command tracking is tackled next chapter.
5.2. Synthesis of Stabilizing Control Law based on the Acceleration Sensitivity
Function (ASF) 117
5.2.2 Control Law Gains Structure

From what precedes, we know that whereas the short-period mode – represented by two eigenvalues–
is easily controlled by feeding back (α, q) or (nz , q) 1 , the integrator pole can only be controlled
by a θ feedback.

It should be mentioned that having a pole “stuck” to zero is not necessarily an issue with
classical control design procedures. However H∞ -optimization routines, such as systune which
we use hereafter, only seek stabilizing solutions. An integrator which cannot be controlled using
specified control structure will then lead to a failure of the optimization, hence the need for a θ
feedback2 .

This is indeed a drawback of using the ASF form on the BWB stabilization problem: it
compels the structure of the law, which is exactly what we are trying to avoid by using structured
H∞ synthesis. Moreover a torque disturbance is not particularly characteristic of flight physics.
As expressed in equation (5.3), we are considering wind disturbance on the angle of attack rather
than pitch acceleration q̇. For this reason w disturbance will be used for control synthesis, and
δαw will be used for simulation. Control problem with all feedbacks is displayed on figure 5.1.
Consistently throughout the report, fixed blocks are displayed in green, while tunable blocks are
displayed in orange.

Finally, as already mentioned in section 5.1.1.3, flight control law gains output is given in
terms of an equivalent elevator deflection δmequi :

δmequi = Kq q + Knz δnz + Kθ δθ (5.36)

This equivalent order δmequi , which would correspond to a single elevator deflection on a
conventional airplane, needs to be allocated to all trailing-edge control surfaces through a control
allocation module. Literature review on control allocation techniques is provided in section 1.1.3.3.
In this study δmequi is linearly allocated to each control surface δmi through the allocation
matrix Kalloc defined in equation (5.14).

Each control surface position is then computed as:

i
δmi = Km δmequi (5.37)

A main originality of our work is to consider Kalloc not as a fixed allocation but a tunable
one: Kalloc is also optimized, at least in sections 5.3 and 5.4.

1
similarly to what was done for the inverted pendulum
2
For sake of completeness we should mention that last version systune does include an option for not necessarily
stabilizing the plant. This option was not yet present when our work was performed but could be useful in this
particular case, and could avoid the need for a θ feedback
118 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

5.3 H∞ Synthesis with Optimization of Allocation Gains

After having setup the control problem, we now describe different control synthesis techniques,
and the rationale for each technique. The flight point considered for illustrating the different
control techniques is a low-speed, medium-altitude and light mass flight point:

• M =0.25

• H =8800 ft

• m =200t

This flight point is expected to be a sizing one from a S&C perspective, for it is a low dynamic
pressure case. Yet in sections 5.5 and 5.7 a codesign is performed on the entire flight enveloppe
in order to extensively identify sizing flight points.

Moreover an initial actuators bandwidth of 5Hz is assumed for all actuators. Compared to
classical values ranging from 1.5Hz to 2.5Hz for aircraft ailerons, we chose fast actuators as a
first guess, in order to be certain that the aircraft can be stabilized with this technology. Codesign
of this chapter aims at proving that lower values are acceptable for bandwidth. Moreover a
100ms delay was set, in order to account for delays coming from the acquisition chain: sensors,
discretization... These delays were approximated as second-order Padé filters during linearization.

Finally following values were assumed for reference model:

ω0 = 1.5 rad.s−1 (5.38)


ξ0 = 0.7 (5.39)

First synthesis is a pure H∞ minimization. We simply seek the minimal value of the H∞ norm
of the transfer function from disturbance w to weighted output zinf , using control laws and
allocation as variables, with the constraint of internally stabilizing the aircraft. Formally speaking
the problem reads:

min kTw→zinf (P, Kq , Knz , Kθ , Kalloc )k∞ (5.40)


(Kq ,Knz ,Kθ )∈R3 ,Kalloc ∈R5

such that: [Kq , Knz , Kθ ] internally stabilizes P

systune routine is used to solve this problem, with plant model depicted on figure 5.1. Optimal
H∞ value obtained is:
γ∞ = 1.289 (5.41)
This value is consistent with Bode magnitude response of the ASF presented on figure 5.4.

The obtained control law is then evaluated on a temporal response to an initial condition
5.3. H∞ Synthesis with Optimization of Allocation Gains 119

12 Control Surfaces Response


δm1
Elevons Deflections (˚)
10 δm2
δm3
8 δm4
δm5
6
4
2
0
-2
-4
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
60 Control Surfaces Deflection Rates
˙ 1
δm
˙ 2
Deflection Rates (˚/s)

40 δm
˙ 3
δm
20 ˙ 4
δm
˙ 5
δm
0

-20

-40

-60

-80
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

2.5 Aircraft Response


δα
2 q
δθ
Parameters (in ˚and ˚/s)

1.5 δnz
1

0.5

-0.5

-1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

Figure 5.2: Temporal simulation in response to an initial condition α0 on angle of attack, with
control law computed through H∞ synthesis.
120 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

10
Kalloc ∗ Kq
9 mδmi /mδm1
8

δmL
5 δmL
4 δmL
3 δmL
2 δmL
1 δmR
1 δmR
2 δmR
3 δmR
4 δmR
5

Figure 5.3: Control allocation gains obtained with H∞ synthesis, and control surfaces pitch
effectiveness. mδmi is the i-th elevon pitch evectiveness defined in equation (5.34).

Maximum gain as a function of frequency


2.2 dB at 9.95 rad.s−1
Singular Values (dB)

Frequency (rad/s)

Figure 5.4: Bode Magnitude response of the ASF after H∞ optimization.


5.4. H2 /H∞ Synthesis with Optimization of Allocation Gains 121

α0 = 2˚ on the angle of attack. Resulting curves are displayed on figure 5.2. Clearly convergence
of aircraft outputs towards zero is obtained at the expense of high control amplitude and rate. In
particular optimal allocation from a H∞ perspective is not physically acceptable: only one control
surface, namely elevon 3, is used, while the others remain at rest.

This observation is confirmed by control allocation gains displayed on figure 5.3. On this figure
Kalloc is represented multiplied by Kq : indeed from one optimization to another the allocation
gains magnitude may vary, but control law gains then vary accordingly. Only the product of the
two remains unchanged.

A similar remark holds for gains values displayed on table 5.1. These values are represented
multiplied by the maximal value of Kalloc in order to be able to make comparisons between
different syntheses.

Kq (sd) Knz (rad) Kθ (sd)


H∞ 8.911 -0.256 5.137
H2 /H∞ 0.976 0.084 0.757

Table 5.1: Gains values for different control synthesis techniques. All values are shown multiplied
by max Kalloc for each synthesis, in order to make fair comparisons.

As a conclusion for pure H∞ synthesis, this optimization procedure has the advantage of
providing the minimal theoretical value for the H∞ criterion γ∞ . It could be easily shown (see
for instance [Fezans et al., 2007; Alazard, 2013]) that ASF based H∞ design could provide a
performance index γ∞ = 1 in the case perfect actuators (i.e. infinite bandwidth), or in the case
of a full-order controller. The gap between the optimal value γ∞ = 1.289 and 1 is due to the
limited actuators bandwidth (5Hz), delays and the static output feedback. This gap is minimized
by the optimization routine.

Yet achieving this optimum is at the expense of large control amplitude, which is to be avoided.
As a consequence a method has to be setup, which provides a similar level of performance, with
reduced control amplitude.

5.4 H2 /H∞ Synthesis with Optimization of Allocation Gains

From the identified drawbacks of pure H∞ minimization, a second strategy is proposed: we seek
minimizing actuators deflection rate, while preserving an adequate closed-loop performance. More
precisely the optimization criterion of this synthesis is the H2 norm of transfer from w to Uad :
we want to minimize the impact of an input disturbance w to the control surfaces deflection rate,
deflection rate being directly proportional to the actuators power consumption. Let us remind that
the H2 norm may be interpreted:

• as the RMS value of the output –in our case the deflection rates vector Uad – in response
to a unitary power white noise input – here on pitch acceleration w .
122 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

Control Surfaces Response


Elevons Deflections (˚) 3 δm1
δm2
2.5 δm3
δm4
2 δm5

1.5

0.5

-0.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
35 Control Surfaces Deflection Rates
˙ 1
δm
Deflection Rates (˚/s)

30 ˙ 2
δm
˙ 3
δm
25 ˙ 4
δm
˙ 5
δm
20
15
10
5
0
-5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

2.5 Aircraft Response


δα
2 q
δθ
Parameters (in ˚and ˚/s)

1.5 δnz
1

0.5

-0.5

-1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

Figure 5.5: Temporal simulation in response to an initial condition α0 on angle of attack, with
control law computed through H2 /H∞ synthesis.
5.4. H2 /H∞ Synthesis with Optimization of Allocation Gains 123

2
Kalloc ∗ Kq
1.8 Kalloc /Kalloc1
1.6 mδmi /mδm1

1.4
1.2
1
0.8
0.6
0.4
0.2
0
δmL
5 δmL
4 δmL
3 δmL
2 δmL
1 δmR
1 δmR
2 δmR
3 δmR
4 δmR
5

Figure 5.6: Control allocation gains obtained with H2 /H∞ synthesis, and control surfaces pitch
effectiveness.

Maximum gain as a function of frequency


2.57dB at 4.25 rad.s−1
Singular Values (dB)

Frequency (rad/s)

Figure 5.7: Bode Magnitude response of the ASF after H2 /H∞ optimization.
124 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

• in temporal domain, as the L − 2 norm of the outputs to an impulse input.

Yet we also want good performance on disturbance rejection: this is expressed through a
constraint on the H∞ norm of the ASF . More precisely we allow a slight increase compared to
the optimal value γ∞ , given by relaxation factor k ≥ 1. In practice we chose k = 1.05, that is
an increase of 5% over optimal H∞ value γ∞ is allowed. The H2 /H∞ problem formally reads:

min kTw→Uad (P, Kq , Knz , Kθ , Kalloc )k2 (5.42)


(Kq ,Knz ,Kθ )∈R3 ,Kalloc ∈R5

such that: Tw→zinf (P, Kq , Knz , Kθ , Kalloc ) ≤ k × γ∞ (5.43)



[Kq , Knz , Kθ , Kalloc ] internally stabilizes P

Once again optimization is run with systune routine. Following optimal values are obtained:

hopt opt
2 = kTw→Uad k2
= 11.89 (5.44)
Tw→zinf (P, Kq , Knz , Kθ , Kalloc ) = 1.342 (5.45)

Temporal simulations for an initial condition on angle of attack are displayed on figure 5.5.
Clearly control deflections are much smaller than for the pure H∞ minimization, with a peak
response at 30˚/s for elevon 2 at the initial time. A closer look at temporal responses reveals a
“smart” control allocation, in the sense that most effective control surfaces in pitch, namely elevon
2, then 5, then 4... are most used, in order to limit deflection rates. This observation is confirmed
by the analysis of figure 5.6. In blue Kalloc × Kq is represented, in order to enable comparison of
control allocation gains with figure 5.3. Allocation gains are decreased by an order of magnitude,
justifying the process of minimzing the H2 norm of deflection rates. Then allocation gains are also
plotted together with elevons pitch rate effectiveness mδmi , all normalized by values for elevon
1. An exact correlation is found between those two quantities: the more the control surface is
effective in pitch, the more it is used for control by the allocation. More precisely pseudo-inverse
allocation, which is a classical solution of control allocation, is found by the optimizer.

Linear control allocation problem is to find a control vector u such that Bu = M , M


being a vector of pseudo-moments. For instance these are control moments in pitch, yaw and roll
computed by the flight control system, when there are more effectors – that is dimension of u–
than moments to create – dimension of M . B is the effectiveness matrix, converting controls
into moments.

Among all solutions, and provided B is full-rank, the pseudo-inverse solution u+ = B T (BB T )−1 M
5.5. Codesign of Control Law Gains and Actuators Bandwidth with Cost Function on
Deflection Rate 125
verifies:

u+ = min kuk2 (5.46)


such that: Bu = M (5.47)

Pseudo-inverse allocation is the minimum norm solution – in the sense of the 2-norm – to the
control allocation problem.

In our problem B restricts to a line matrix Bq = [. . . mδmi . . . ], u = [. . . δmi . . . ] and


M = mδmequi δmequi . We can assume mδmequi = 1 without loss of generality. Then the pseudo-
inverse matrix Bq+ reads:

 .. 
.
δmi
 
Bq+ =  P10 (5.48)
 
δm 2
 i=1 i 
..
.

To ensure that the allocation has found the pseudo-inverse solution, element-wise division of
Kalloc and Bq+ gives:
−1.8665
 
−1.8835
 
−1.8967
 
 
−1.8701
 
−1.8563
+
Kalloc ./Bq = 
 
−1.8665

 
−1.8835
 
−1.8967
 
 
−1.8701
−1.8563

We now know that the H2 /H∞ synthesis successfully recaptures the analytical solution
of the control allocation problem. Since this pseudo-inverse solution provides a satisfac-
tory behavior from a control amplitude perspective, we choose to keep this solution for
the rest of the chapter. As a consequence, from now on Kalloc is no more a tunable
matrix but is chosen equal to the pseudo-inverse Bq+ . It has the desirable side-effect of
reducing the number of optimization variables, hence enabling a more efficient codesign
procedure in following sections.
126 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

q √
3Lz
ew 2Lz 1+ Ve s wz −zα
σz πVe (1+ Lz s)2 Ve
Ve

Turbulence Model
w s2 +2ξω0 s+ω02
zinf
s2

δαw
+ α
δ α̇ + δα
q
+
Ua Ẋa q̇ 1 q Xa Ya θ
Ba + +
s Ca +
nz
+ δ θ̇ δθ +

Aa
Da
Longitudinal Aircraft Model

+ Kq Delay
1 Uad u
s Kwact +
Kalloc δmequi
+ Kθ Delay

+ Knz Delay

Actuators Control Law


P (s)

WHM × Uad
WHM

Figure 5.8: Block-diagram of stabilizing control law and actuators bandwidth optimization com-
puted with ASF scheme. Fixed parameters are in green, and tunable parameters are in orange.

5.5 Codesign of Control Law Gains and Actuators Bandwidth with


Cost Function on Deflection Rate

Now that a relevant synthesis formulation has been setup, this same H2 /H∞ formulation is
applied for codesign of control law gains and actuators bandwidth. The objective is to improve
previously found optimal value hopt2 , with degrees of freedom not only on control gains, but
5.5. Codesign of Control Law Gains and Actuators Bandwidth with Cost Function on
Deflection Rate 127

3 Control Surfaces Response


δm1
2.5 δm2
Elevons Deflections (˚)

δm3
2 δm4
δm5
1.5
1
0.5
0
-0.5
-1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
20 Control Surfaces Deflection Rates
˙ 1
δm
˙ 2
δm
15
Deflection Rates (˚/s)

˙ 3
δm
˙ 4
δm
˙ 5
δm
10

-5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

2.5 Aircraft Response


δα
2 q
δθ
1.5 δnz
Parameters (in ˚and ˚/s)

0.5

-0.5

-1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

Figure 5.9: Temporal simulation in response to an initial condition α0 on angle of attack, with
control law and actuators bandwidth computed through codesign with objective set on kTw→Uad k2
.
128 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

4
HM /HM 1
3.5 ωact
mδmi /mδm1

2.5

1.5

0.5

0
δmL
5 δmL
4 δmL
3 δmL
2 δmL
1 δmR
1 δmR
2 δmR
3 δmR
4 δmR
5

Figure 5.10: Optimized actuators bandwidth in red, plotted together with normalized hinge mo-
ments in blue, and normalized elevons pitch effectiveness, in green.

Maximum gain as a function of frequency

2.54dB at 3.84 rad.s−1


Singular Values (dB)

Frequency (rad/s)

Figure 5.11: Bode Magnitude response of the ASF after H2 /H∞ codesign on actuators band-
width.
5.5. Codesign of Control Law Gains and Actuators Bandwidth with Cost Function on
Deflection Rate 129
also on actuators bandwidth. Let us remind that previous syntheses were performed with 5Hz
actuators on all control surfaces. Kalloc being now fixed to the pseudo-inverse solution, the
problem formally reads:

min Tw→Uad (P (Kwact ), Kq , Knz , Kθ ) (5.49)


(Kq ,Knz ,Kθ )∈R3 ,Kwact ∈R5 2

such that: Tw→zinf (P (Kwact ), Kq , Knz , Kθ ) ≤ k × γ∞ (5.50)



[Kq , Knz , Kθ ] internally stabilizes P (Kwact )

with:

ωact 1
 
... 0
 . .. ..
 ..

 . . 

 . ..
 ..

ωact 5 . 
Kwact = . (5.51)
 
 . .. 
 . ωact 1 .


 .. .. ..
 
 . . .


0 ... ωact 5

Assuming symmetrical elevons should have the same actuators, Kwact has five degrees of
freedom. The closed-loop model is then depicted on figure 5.8.

After optimization following values were obtained:

kTw→Uad kopt
2 = 11.5217 (5.52)
kTw→zinf k∞ = 1.340 (5.53)

Value 5.52 should be compared to value 5.44: introducing Kwact as a design variable entails
a 3% improvement on the H2 -norm criterion. This may not seem a lot, but a comparison of
figures 5.5 and 5.9 provides a much clearer view on the improvements allowed by minimizing
˙ is decreased by a factor 2,
actuators bandwidth: initial peak response of deflection rates δm
from 30˚/s to 15˚/s. This is obtained at the expense of few degradation on states response, for
the constraint on the H∞ norm of the ASF ensures a maximal 5% degradation over γ∞ .
Template and closed-loop ASF are shown on figure 5.11.

Codesign results on actuators optimal bandwidth are displayed on figure 5.10. Minimal de-
˙ are ensured by using in priority most effective control surfaces, namely 2, 4
flection rates δm
and 5. An interesting result is that “faster” bandwidth remain at reasonable levels. Maximum
optimized bandwidth obtained is slightly above 3Hz, which is within current technology capabili-
130 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

ties. Of course results are presented here only for a single flight point; sizing results on the whole
flight domains are shown hereafter. Yet this already gives confidence in the capabilities of the
configuration to be correctly stabilized.

However a last issue in the codesign problem formulation remains: consistently with the
objective defined in equation (5.49), in order to minimize deflection rates, most effective control
surfaces are allocated a faster bandwidth. Yet these most effective elevons have also the highest
area, and possibly highest hinge moments. As a consequence high power consumption may result
from this optimization. Therefore we propose a last improvement of the codesign process: instead
of only minimizing the H2 -norm of Uad in response to w , we will now weight Uad by
respective control surface maximum hinge moment. Consequently the optimization criterion will
now be directly a monotonous function of the actuators mechanical power consumption.

5.6 Evaluation of Maximum Aerodynamic Hinge Moments

For conciseness purpose, next chapter is entirely dedicated to the computation of aerodynamic
coefficients for the BWB and derived control surfaces layouts. Hence we chose to describe
the way maximum aerodynamic hinge moments are computed only in next chapter 6. Here we
only give results of this process in terms of maximum hinge moment as a function of flight case
displayed on figure 5.12. These maximum hinge moments are used in this section to weight the
control signal Uad .

Let us just point out that these hinge moments were computed using conservative hypotheses:
a vortex-lattice method was used for evaluating these hinge moments, and it is well known that
these methods over-estimate the aerodynamic effects for large control surfaces deflections, by
assuming linear aerodynamics. Calibration factors used for our study come from lift coefficient
calibration only: no appropriate data was available concerning hinge moments magnitude for a
BWB . Assuming conservative hypotheses was a computation choice for sizing. As a consequence
hinge moments values displayed here are for sure largely over-estimating the actual hinge moments,
even though large HM are often pointed out as a threat for BWB design, as shown in section 1.1.
The only other HM values for BWB configuration we managed to find in literature come from
[Garmendia et al., 2015b]; smaller yet of similar order of magnitude HM values are found, with
a similar aerodynamic method, for a smaller aircraft configuration.

Nevertheless we believe the method we propose here is still valid, but would deserve being
applied again with HM values coming from higher-fidelity aerodynamics. We therefore advise
readers to focus on the method developed rather than numerical results.

5.7 Codesign of Control Law Gains and Actuators Bandwidth with


Cost Function on Actuators Power Consumption

For codesign weighted by hinge moments, the optimization problem formally becomes:
5.7. Codesign of Control Law Gains and Actuators Bandwidth with Cost Function on
Actuators Power Consumption 131

(a) Elevon 1. (b) Elevon 2.

(c) Elevon 3. (d) Elevon 4.

(e) Elevon 5.

Figure 5.12: Maximum hinge moment, computed for δmi = 30˚, as a function of flight point for
each elevon.

min kTw→WHM ×Uad (P (Kwact ), Kq , Knz , Kθ )k2 (5.54)


(Kq ,Knz ,Kθ )∈R3 ,Kwact ∈R5

such that: kTw→zinf (P (Kwact ), Kq , Knz , Kθ )k∞ ≤ k × γ∞ (5.55)


[Kq , Knz , Kθ ] internally stabilizes P (Kwact )
132 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

Control Surfaces Response


8
δm1
δm2
Elevons Deflections (˚)

6 δm3
δm4
4 δm5

-2

-4
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
25 Control Surfaces Deflection Rates
˙ 1
δm
20 ˙ 2
δm
Deflection Rates (˚/s)

˙ 3
δm
15 ˙ 4
δm
˙ 5
δm
10

-5

-10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

2.5 Aircraft Response


δα
2 q
δθ
1.5 δnz
Parameters (in ˚and ˚/s)

0.5

-0.5

-1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

Figure 5.13: Temporal simulation in response to an initial condition α0 on angle of attack,


with control law and actuators bandwidth computed through codesign with objective set on
kTw→WHM ×Uad k2 .
5.7. Codesign of Control Law Gains and Actuators Bandwidth with Cost Function on
Actuators Power Consumption 133
4
HM /HM 1
ωact , (Hz)
3.5 mδmi /mδm1

2.5

1.5

0.5

0
δmL
5 δmL
4 δmL
3 δmL
2 δmL
1 δmR
1 δmR
2 δmR
3 δmR
4 δmR
5

Figure 5.14: Optimized actuators bandwidth in red, plotted together with normalized hinge mo-
ments in blue, and normalized elevons pitch effectiveness, in green. Values are obtained after an
optimization with a minimization objective on kTw→WHM ×Uad k2 .

Maximum gain as a function of frequency


2.54 dB at 4.02 rad.s−1
Singular Values (dB)

Frequency (rad/s)

Figure 5.15: Bode Magnitude response of the ASF after H2 /H∞ codesign on actuators band-
width with weighting on the actuators power consumption.
134 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

WHM is a diagonal matrix gathering hinge moments for all control surfaces:

WHM = diag(HMi , i = 1 . . . 10) (5.56)

WHM is also displayed on figure 5.8.

Remark 5.3
Following definition of the L2 -norm for the MIMO case in equation (3.24), we note that the unit
of kTw→HM ×Uad k2 is not directly Watts but W.s3/2 , because of the frequency integration. As a
consequence the minimization criterion is not exactly power consumption. However this criterion
is monotonous in power consumption HM × δm. ˙

After optimization following values are found:

kTw→WHM ×Uad kopt


2 = 2.5072.10
5
(5.57)
kTw→zinf k∞ = 1.340 (5.58)

Bandwidth allocation strategy, shown on figure 5.14, is quite different from previous opti-
mization: in order to minimize power consumption, only small control surfaces (in terms of area,
then also in terms of hinge moment) are used as fast, stabilizing control surfaces. More precisely
for this flight point only elevon 5 has a significant bandwidth, even though it remains largely
acceptable: 2.3Hz. On temporal simulation of figure 5.13, we see that satisfactory stabilization is
obtained at the expense of larger gains, implying larger control deflections: maximum deflection
is 8˚ versus 3˚ with previous optimization, and maximum deflection rate is 20˚/s versus 15˚/s. Yet
these maximum values are obtained for elevon 5 versus elevon 2 for previous codesign: and it is
far more reasonable to move a 6.5m2 control surface at 20˚/s than a 38.5m2 control surface at
15˚/s (see table 2.3 for values of control surfaces areas). Increased gains magnitude is confirmed
by comparison of feedback gains found by the two optimizations, in table 5.2.

Kq (sd) Knz (rad) Kθ (sd)


codesign 1.3815 0.0668 0.6576
codesign with weighting by WHM 5.8604 0.0492 1.5549

Table 5.2: Gains values computed by codesign optimization, without and with weighing by hinge
moments. All values are shown multiplied by max Kalloc for each synthesis, in order to make fair
comparisons.

As a preliminary conclusion, on this flight point our goal is achieved: we found a design solution
that properly stabilizes the aircraft, while minimizing the impact on actuators power consumption.
5.8. Actuators Bandwidth Sizing on the Whole Flight Envelope 135

Actuators Bandwidth (Hz) 3

2.5

1.5

0.5

0
δm1 δm2 δm3 δm4 δm5

Figure 5.16: Actuators optimal bandwidth after codesign with hinge moments weighting, for all
flight points.

5.8 Actuators Bandwidth Sizing on the Whole Flight Envelope

Now that the methodology has been setup, we propose to apply it on the whole flight domain
described in section 2.3. For each flight point, an aerodynamic model with hinge moments is
computed, and actuators bandwidth are sized with the codesign methodology presented in this
section.

Results of optimal bandwidth are displayed on a bar graph on figure 5.16. Consistently with
previous section, nearly only elevon 5 is used with a high dynamics for stabilization. Real good
news is, for no flight point does the optimized bandwidth of elevon 5 go beyond 2.6Hz. This
means with quite classical actuators, the instability can be controlled. Yet readers should not be
mistaken: we still have no guarantee of not overshooting physical stops.

Here we provide a deeper insight into the optimum values of optimized bandwidth of elevon
5. Results as a function of flight point are displayed for each mass on figure 5.17. The main
interesting fact of these results is that requirements for the bandwidth are higher in “corner” flight
points combining high altitude and low speed. For these points the dynamic pressure is lower,
hence control surfaces are less aerodynamically efficient. Moreover as it has been already seen,
these flight points correspond to the maximum instability level. The optimizer then retrieves such
physical considerations.
136 Chapter 5. Longitudinal Codesign through H2 /H∞ Synthesis

Conclusion

In this chapter the interest of considering optimization methods for closed-loop physical sizing was
demonstrated: we now have an order of magnitude of minimal power consumption and associated
physical bandwidth requirements in order to properly stabilize the aircraft. Sizing flight points
are also identified by the process: they correspond to low-speed cases, which are traditionally
sizing from a handling qualities perspective. We have also demonstrated that the aircraft can
be stabilized longitudinally with quite common actuators bandwidth, which up to now remained
questionable. A frequency segregation between larger control surfaces aimed at a slow motion,
and smaller control surfaces for high bandwidth stabilization, was obtained. From a control theory
perspective, we found that the nonsmooth optimization synthesis was able to find numerically the
analytical solution of the control allocation problem. For this reason from now on the analytical
pseudo-inverse solution will be used instead of being numerically optimized. Limits of the ASF
scheme were also evidenced: in particular putting the plant dynamics under the form of a second-
order differential equation imposes constraints on the feedback structure, which we are precisely
trying to avoid by using synthesis tools for structured compensators. Moreover the ASF form was
still applicable for longitudinal dynamics, but will be more difficult to apply on lateral dynamics.
For these reasons a new control synthesis scheme will be applied in next chapters. Moreover some
points still lack in our study: control surfaces sizing is influenced not only by longitudinal control,
but also by lateral control, which until now was not considered. As a consequence next studies will
include a three-axes flight control laws synthesis. Then the codesign procedure we set up does not
provide a certificate that control magnitude for stabilizing the aircraft is admissible: deflections
and rates may exceed physical limits. The optimization criterion is also not straightforward for
physical interpretation: even though we know it is monotonous as a function of power consumption,
numerical optimal value is hard to interpret. Finally, more than actuators bandwidth, we would
like to size the control surfaces area needed for control. All these remarks will then lead to the
study presented in next chapters.
5.8. Actuators Bandwidth Sizing on the Whole Flight Envelope 137

(a) Optimized bandwidth for elevon 5 as a function of (b) Optimal value of criterion kTw→HM ×Uad kopt
2 as
flight point, for m = 200t. function of flight point, for m = 200t.

(c) Optimized bandwidth for elevon 5 as a function of (d) Optimal value of criterion kTw→HM ×Uad kopt
2 as
flight point, for m = 250t. function of flight point, for m = 250t.

(e) Optimized bandwidth for elevon 5 as a function of (f) Optimal value of criterion kTw→HM ×Uad kopt
2 as
flight point, for m = 300t. function of flight point, for m = 300t.

(g) Optimized bandwidth for elevon 5 as a function of (h) Optimal value of criterion kTw→HM ×Uad kopt
2 as
flight point, for m = 350t. function of flight point, for m = 350t.

Figure 5.17: Maximum hinge moment, computed for δmi = 30˚, as a function of flight point for
each elevon.
Chapter 6

Three-Axes Aerodynamic Model for


Handling Qualities Parameterized by
Control Surfaces Size

This chapter is dedicated to obtaining a state-space representation of the aircraft dynamics,


which is parameterized by the span of trailing-edge elevons. For that purpose, a geometrical
model of the BWB parameterized by the relative span of trailing-edge control surfaces is
developed. Then aerodynamic effectiveness of control surfaces is computed both in longitu-
dinal and lateral. More precisely the aerodynamic gradients are computed using light CFD
methods, and then calibrated using more accurate data. These data come from an aerody-
namic model for handling qualities of the initial configuration, coming from previous CFD
and wind tunnel studies. Next step consists in approximating in a polynomial way all the
state-space representations obtained for different control surfaces spans, knowing that in the
end a continuously parameterized LFR is necessary for a further optimization of this span.
Our main contributions in this chapter are:

• Developing an method for aerodynamic computation combining fast data generation


and relative accuracy, through calibration with an existing aerodynamic model.

• Using the formalism of Linear Fractional Representation for an original application, i.e.
for obtaining a parameterized representation of aerodynamic models with a varying
geometry.

Contents
6.1 Parameterized Geometrical Model of Blended Wing-Body Control Sur-
faces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
6.2 Computation of Control Surfaces Aerodynamic Efficiency for Several
Geometrical Configurations . . . . . . . . . . . . . . . . . . . . . . . . 141
6.2.1 Computation of Lift Coefficient . . . . . . . . . . . . . . . . . . . . 142
6.2.2 Computation of Pitching Moment Coefficient . . . . . . . . . . . . . 144
6.2.3 Computation of Rolling Moment Coefficient . . . . . . . . . . . . . . 144
6.3 LFR Approximation of State-Space Representations . . . . . . . . . . . 145
6.4 Computation of Maximum Hinge Moments for Evaluation of Power
Consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

139
Chapter 6. Three-Axes Aerodynamic Model for Handling Qualities Parameterized by
140 Control Surfaces Size

From the conclusions of previous chapter, we know now that the codesign approach is promising
for designing some physical parameters directly taking into account some closed-loop constraints.
However the problem formulation used up to now revealed some drawbacks, already addressed in
the conclusion of previous chapter. Taking these remarks into account, we now want to reuse the
codesign approach, on a more concrete problem from an aircraft design perspective:

• We want to directly size the elevons, instead of just their bandwidth.

• Elevons sizing being affected by both longitudinal and lateral-directional criteria, we need
to take all these criteria into account in the design.

Yet in order to feed the optimizer, a state-space representation of the aircraft dynamics,
continuously parameterized by a parameter representing the elevons size, needs to be developed.
This is presented in current chapter. More precisely discrete aerodynamic models are computed for
each flight point and elevon size, then a continuous approximation of these approximated models
is obtained. The ultimate goal of this part is that the continuous optimizer used in chapter 7 may
use the continuous variable representing elevons size as an optimization variable.

6.1 Parameterized Geometrical Model of Blended Wing-Body Con-


trol Surfaces

(a) η = 1 (b) η = 0.8 (c) η = 0.4

Figure 6.1: Elevons size for different values of parameter η.

In “classical” plant-controller optimization problems applied to Aeronautics, attention is drawn


on minimizing elevator size under longitudinal constraints. The BWB lacking by definition an
elevator, this process is not applicable here. Therefore outer elevons total span is chosen as a
plant figure-of-merit to be minimized. More precisely a variable η representing the ratio of outer
elevons total span compared to initial control surfaces span is introduced. On figure 6.1(a) initial
elevons layout, corresponding to parameter value η = 1 is shown. As presented in section 2.2, this
initial layout features five control surfaces on each side of the wing spanning the whole trailing
edge, except a gap between elevon 1 and 2 for engine pylon integration – elevons are numbered
from inboard to outboard. Elevons relative chord is limited in x−wise position by cabin integration
for elevon 1, and by rear spar for elevons 2 to 5. So we decided to keep elevons relative chord
constant, as well as elevons number – as mentioned in chapter 4, elevons number is mostly a
6.2. Computation of Control Surfaces Aerodynamic Efficiency for Several Geometrical
Configurations 141
failure case problem that is out of scope of our study. Also initial elevons are split equally on the
outer wing, that is elevons 2 to 5 have the same span.

From this initial layout, control surfaces were derived using following assumptions:

y2inb
• η= y2init
, y2inb and y2init
inb
being derived and initial inboard y-position of elevon 2 respectively.
inb

• y5out = y5init
out
, y5out and y5init
out
being derived and initial inboard y-position of elevon 5 respec-
tively.

• Elevons relative chords are kept constant and equal to 22%.

• Elevons 2 to 5 are split equally.

• Elevon 1 is kept constant.

It is then clear that 0 ≤ η ≤ 1, η = 1 corresponding to initial elevons layout and η = 0


corresponding to the lack of any control surface on the outer wing. Examples of layouts for
η = 0.8 and η = 0.4 are presented on figures figure 6.1(b) and figure 6.1(c) respectively.

6.2 Computation of Control Surfaces Aerodynamic Efficiency for


Several Geometrical Configurations

Reference Aircraft
Aerodynamic Model

Aircraft with varying AVL APRICOT


elevons span Aerodynamic Calibrated Aerodynamic Toolbox
Geometrical Model Computation Models LFR Approximation
η = 0.1 : 1
η = 0.1 : 1
mass Mach H mass Mach H

LFR function of
elevons span

mass Mach H

Figure 6.2: Process for obtaining LFR approximation of state-space representations as a function
of η parameter.
Chapter 6. Three-Axes Aerodynamic Model for Handling Qualities Parameterized by
142 Control Surfaces Size
0.5

0.4 0.15

0.3
Czδmi (1/rad)

Czδmi (1/rad)
0.10
0.2

0.1 0.05

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Mach Mach
elevon 1 2 3 4 5 elevon 1 2 3 4 5
(a) Comparison of lift gradient coefficients for refer- (b) Comparison of lift gradient coefficients for project
ence aircraft: CLref
δm
(dotted) vs CLAVδm
L
(plain). aircraft before and after calibration respectively:
i i
Lproj
CLAV
δmi
(plain) vs CLproj
δmi
(dotted) Projet aircraft
here is for η = 0.6.

Figure 6.3: Comparison of pitch gradient coefficients CLδmi from AVL outputs with reference
aircraft aerodynamic data (left) and with calibrated data (right).

Computation of aerodynamic model for different parameter values of η is described in this


section. For sake of clarity initial layout – namely configuration for η = 1 – is called “reference”
aircraft, whereas derived layouts – namely configurations for η < 1 – are called “project” aircraft.
Reference flight dynamics model described in chapter 2 is used for all configurations described in
section 6.1 as planform and airfoils are kept constant for all configurations; therefore only control
surfaces aerodynamic effectiveness needs to be evaluated for the project aircraft. Once again the
goal of this chapter is to obtain state-space representations continuously parametrized by elevon
span parameter η . Process to obtain such a continuous approximation is presented on figure 6.2.
A first step is to compute calibrated aerodynamic models for discrete values of η , namely for
values between 0.1 and 1 with steps of 0.1. For that purpose the Athena Vortex-Lattice (AVL)
software [Drela and Youngren, 2006] was used together with calibration factors coming from the
supposedly known aerodynamic coefficients of the initial BWB design. This allows for taking into
account Mach effects and non-linearities due to high angle of attack or high control deflection,
which AVL cannot account for. Continuous approximation of discretized models is treated in
section 6.3, whereas computation of calibrated aerodynamic models is explained hereafter.

6.2.1 Computation of Lift Coefficient

The aircraft total lift coefficient Cz comprises a i-th control surface deflection dependency Cz i
which can be written as:
NL
Cz i = kδmi
(α, δmi )Czδmi δmi (6.1)

N L (α, δm ) accounts for loss of control surface effectiveness as a function of the angle
where kδm i i
6.2. Computation of Control Surfaces Aerodynamic Efficiency for Several Geometrical
Configurations 143
of attack α and control surface deflection δmi , and Czδmi is the lift gradient of the i-th control
N L (α, δm ) is known for reference aircraft, and kept for project aircraft. Reference lift
surface. kδm i i
gradient Czδmi ref is supposed known from previous studies [Meheut et al., 2012], and is compared
with the lift gradient computed by AVL for the reference configuration Czδmi AV L (see figure
6.3(a)). From these data a calibration factor accounting for AVL lack of accuracy is computed:

Czδmi ref
∆Cz AV
i
L
= (6.2)
Czδmi AV L

Then an AVL computation is run on project geometry in order to compute the lift gradient
proj
of project aircraft Czδmi AV L . This gradient is finally calibrated using the previously computed
calibration factor ∆Cz AV
i
L of equation (B.2) (see figure 6.3(b)), giving:

proj
Cz proj
i
NL
= kδm i
∆Cz AV
i
L
Czδmi AV L δmi (6.3)

As its name suggests, AVL is a vortex lattice method; lifting surfaces are treated as panels.
The meshing of the BWB is visible on figure 6.4. An equal number of x-wise cells is chosen
throughout the span: each local chord is decomposed into 45 panels. This allows for a finer
meshing of the outer wing, which is the main region of interest, compared to the center body.
Then the chord-wise meshing is not equally spaced along the chord: rather it follows a cosine
distribution. This means that leading and trailing edges are more finely meshed, so linear gradients
of the control surfaces, which is the main objective of this part, are better captured. Yet despite
these precautions it remains a linear aerodynamic method, with all the limitations it implies. Its
speed makes it well adapted to handling qualities evaluation at future projects level, but non-linear
effects cannot be captured. This is the reason why the previously explained calibration method is
developed.

Figure 6.4: Panels distribution for the BWB lifting surfaces.


Chapter 6. Three-Axes Aerodynamic Model for Handling Qualities Parameterized by
144 Control Surfaces Size
6.2.2 Computation of Pitching Moment Coefficient

A similar process is used for computing control surfaces pitching moment effectiveness of project
configurations. More precisely the aircraft pitching moment coefficient Cm comprises a i-th
control surface deflection dependency Cmi which can be written as:

NL
Cmi = kδmi
(α, δmi )Czδmi δmi (XCG − XFi ) (6.4)

where XCG denotes the x−wise CG position, and XFi denotes the i-th control surface
aerodynamic center x−wise position. However XFi is not a direct output of AVL and has to be
computed as follows:
AV L
Cm δm
XFi AV L = XCG − AV Li (6.5)
Czδm
i

This leads to computing the aerodynamic center calibration factor for the reference aircraft
knowing the actual aerodynamic center XFi ref for the i-th elevon:

XFi ref
∆XFi AV L = (6.6)
XFi AV L

Finally an AVL computation is run on project aircraft in order to compute the pitching
moment gradient of project aircraft i-th elevon CmG AV Lproj , and the resulting aerodynamic
δmi
center is computed and calibrated as follows:
proj
 
AV L
Cm
Lproj δm
XFi AV = ∆XFi AV L . XCG − i
Lproj
 (6.7)
CzAV
δm i

6.2.3 Computation of Rolling Moment Coefficient

A similar process is used for computing control surfaces rolling moment effectiveness of project
configurations. More precisely the aircraft rolling moment coefficient Cl comprises a i-th control
surface deflection dependency Cli which can be written as:

NL
Cli = kδmi
(α, δmi )Clδmi δmi (6.8)

Calibration factor for rolling moment coefficient is then computed as follows:

Clδmi ref
∆ClAV
i
L
= (6.9)
Clδmi AV L

And finally rolling moment coefficient of project aircraft is calibrated using calibration factor
6.3. LFR Approximation of State-Space Representations 145

of equation (B.9):
proj
Clproj
i
NL
= kδmi
∆ClAV
i
L
Clδmi AV L δmi (6.10)

This method, summarized on figure 6.2, combines the advantages of fast data generation
through light CFD computation, and far better accuracy than AVL direct output through
accurate knowledge on a reference configuration. Here it is applied only to control surfaces aero-
dynamic coefficients, but the process would be similar for computing any aerodynamic coefficient
of a project aircraft with a baseline knowledge on a similar reference aircraft.

6.3 LFR Approximation of State-Space Representations

Figure 6.5: Linear Fractional Representation.

Once aerodynamic coefficients are computed and calibrated for discretized values of η ,
the final step concerning modeling problem consists in obtaining an approximation of state-space
representations for continuously varying parameter η . For that purpose it was chosen to work with
the LFR framework, because this representation is suited to the optimizer coming from the control
community used in chapter 7. Moreover efficient algorithms for approximating a set of numerical
data as an LFR were developed by ONERA [Roos et al., 2014]. An LFR is a model where
all fixed dynamics are gathered in a single linear time-invariant plant M , whereas uncertainties
or varying parameters are contained in a block-diagonal matrix ∆ (see figure 6.5). Polynomial
and rational expressions are for instance easily convertible into LFR . For a comprehensive LFR
theory, readers may refer to [Zhou et al., 1996].

More precisely the problem is that of finding an LFR approximating as closely as possible
state-space representations computed for different values of η . Uncertainties are not considered
in this study, so the ∆ block is only composed of η parameter repeated several times. Moreover
as mentioned previously, in this work the BWB planform is kept constant, therefore state matrix
A from the state-space representation is assumed to be independent from η . As a consequence
only control matrix B containing elevons effectiveness is approximated by the LFR . However the
described process would be similar if including planform variables having an effect on A matrix.

It was decided to restrict search for LFR approximations to polynomial approximations in


order to keep the LFR order – that is the amount of times η is repeated inside the ∆ block
– as small as possible. From a physical perspective this can be justified by the fact that control
surfaces effectiveness should vary smoothly with respect to their span. The least-squares routine
Chapter 6. Three-Axes Aerodynamic Model for Handling Qualities Parameterized by
146 Control Surfaces Size
lsapprox from APRICOT library on Matlab [Roos et al., 2014] was used. The problem consists
in finding a polynomial P of degree np minimizing following criterion:

N
X
C= [Bηk − P (η)]2 (6.11)
k=1

Bηk being control matrix for sampled values of η , η k,k=1:N = [0.1...1]. Degree of polynomial
P is set to np = 5, and resulting LFR size is 20 – this number represents how many times η
is repeated inside ∆ block. Maximum RMS value is 9.36.10−3 and maximum local absolute
error is 2.01.10−2 . lsapprox instead of orthogonal least-squares olsapprox routine was used
for it achieves higher accuracy. LFR order is however higher but remains acceptable, that
is computation cost for simulation using Simulink [MATLAB, 2014] is only increased by a few
seconds. Resulting approximated aerodynamic gradients are plotted on figure 6.6.

(a) CmLF R
δmi approximation of elevons pitch gradient
LF R
(b) Clδm i
approximation of elevons roll gradient for
for varying η. varying η.

Figure 6.6: LFR polynomial approximations of elevons pitch and roll gradients.

6.4 Computation of Maximum Hinge Moments for Evaluation of


Power Consumption

In section 5.6, hinge moment evaluation was needed in order to evaluate the actuators power
consumption for longitudinal stabilization. How these hinge moments were computed is described
here. These HM were computed only for the reference aircraft, that is for η = 1. Including hinge
moments evaluation into the optimization of control surfaces total span would be an interesting
study, but is left for future work. I-th control surface aerodynamic hinge moment HMi is
computed as follows:

HMi = 1/2ρVair 2 SlCHMi (6.12)

CHMi is the i-th control surface aerodynamic hinge moment coefficient. Contrary to previous
aerodynamic coefficients, hinge moment at δmi = 0˚ is not equal to zero in general, so CHMi is
in first approximation an affine function of δmi , and not a linear function of δmi . AVL only
provides hinge moments coefficients for a given control deflection: two separate computations
6.4. Computation of Maximum Hinge Moments for Evaluation of Power Consumption
147

for two deflections are then necessary in order to be able to compute CHMi for an arbitrary
deflection δmi . We have chosen to compute hinge moments for δmi = 0˚ and δmi = 10˚.

Since we have no value of hinge moments for the reference BWB we have no value to
calibrate with. The rationale is then to use calibration factor obtained for lift gradients, ∆Cz AV
i
L

considering that the error made on the evaluation of Czδmi should be the same order of magnitude
than the error made on the evaluation of CHMi . Of course the accuracy of the calculations
would benefit from higher-fidelity CFD to calibrate on.

Also non-linearities with respect to Mach, angle of attack and control deflection are applied
to the hinge moment coefficient, giving:

CHMi ref (δmi ) = kδm


NL
i
∆Cz AV
i
L
CHMi AV L (δmi ) (6.13)
CHMi AV L (10˚)− CHMi AV L (0˚)
NL
= kδmi
∆Cz AV
i
L
δmi (6.14)
10˚

Finally hinge moment coefficients are evaluated for maximal control deflection δmi = 30˚,
giving:
AV L (10˚) − C AV L (0˚)
AV L CHMi HMi
CHMi ref (30˚) = kδm
NL
i
∆Cz i 30˚ (6.15)
10˚

Resulting hinge moments HMi are displayed on figure 5.12.

Conclusion

In this chapter our method to obtain state-space representations of the aircraft longitudinal and
lateral-directional dynamics parameterized by the control surfaces relative span was developed.
This method combines fast data generation through the use of light CFD software, and reasonable
accuracy through calibration with accurate aerodynamic data. The main advantage of converting
the set of discretized state-space representations into an LFR representation is the ease of use
of this type of objects by control optimization solvers. Now that such a parameterized model is
available, it will be used for a codesign optimization in next chapter.
Chapter 7

Co-Design of Control Surfaces and


Laws under Maneuverability and
Stability Constraints

This chapter aims at conjointly optimizing the control surface sizes and three-axes control
laws under maneuverability constraints and constraints on response to external disturbance.
For that purpose, a Reference Model Tracking (RMT) scheme is developed; first the
RMT principle is introduced on longitudinal synthesis only, and then extended on the three-
axes cases, with different formulations being discussed. A next step consists in translating
handling qualities dynamics constraints into H∞ constraints. Finally the control surfaces total
span, whose parameterization was discussed in previous chapter, is optimized conjointly with
control laws gains. Different formulations are studied: first the control allocation module is
parameterized as an LFR , then it becomes an optimization variable. Differences in resulting
designs are discussed.
Our main contributions to this chapter are:

• Optimizing through a unique H∞ criterion three-axes control laws gains, with a fixed
structure coming from industrial requirements, using a RMT scheme.

• Providing a simple parametrization for control allocation gains as a function of the


control surface size.

• The codesign of control surface span, control laws and possibly control allocation gains
for a closed-loop sizing of the aircraft.

Contents
7.1 Development of a Three-Axes H∞ Model Reference Tracking Scheme 150
7.1.1 Model Reference Tracking for Longitudinal Synthesis . . . . . . . . . 150
7.1.2 Reference Model Tracking for Three-Axes Synthesis . . . . . . . . . 162
7.2 Translation of Handling Qualities Dynamic Constraints into H∞ Con-
straints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.3 Evaluation of Different Methods for Control Surfaces Codesign . . . . 174
7.3.1 Decoupled Optimization with LFR Parametrization of the Allocation 174
7.3.2 Codesign with LFR Parametrization of the Allocation . . . . . . . . . 181

149
150 Chapter 7. Co-Design of Control Surfaces and Laws

7.3.3 Codesign with Tunable Allocation . . . . . . . . . . . . . . . . . . . 188

Chapter 5 concluded the possibility of longitudinal stabilization for the BWB with state-of-
the-art actuators capabilities, yet several limitations of the methods were emphasized. Namely
the ASF scheme for longitudinal H∞ synthesis constrained the control structure (by requiring
a feedback of pitch attitude θ ), which was exactly what we were seeking to avoid. Then
no proof of the maximum amplitude for a given level of disturbance was delivered: rather an
optimal power consumption was guaranteed, which is by itself a result, but does not guarantee
the maximum amplitude of deflections. Finally lateral criteria were not included, whereas elevons
are also necessary for lateral control.

This chapter aims at overcoming all those limitations, using models developed in chapter 6.
More precisely, a new H∞ synthesis scheme is developed: Reference Model Tracking (RMT). Its
main advantage is its easiness to specify a desired dynamics in response to a control input, with
an arbitrary gains structure. The RMT scheme is first developed on a longitudinal case only, and
then extended to the three-axes case, where different formulations are discussed. This is tackled
in section 7.1. Then in order to setup the codesign problem of optimizing control surfaces span
η , as already introduced in previous chapters, constraints on closed-loop handling qualities are
necessary: if no constraint is set on control amplitude, then any disturbance can be counteracted
with an arbitrary small control surfaces area, at the expense of out-of-range control amplitude.
Setting those constraints with the H∞ formalism is detailed in section 7.2. Finally the codesign of
η parameter and control law gains under closed-loop handing qualities constraints is developed
in section 7.3. Three methods are proposed, ranging from the simplest yet least optimal, to the
most optimal yet implying more optimization complexity. The first two methods indeed rely on
a LFR parametrization of the control allocation matrix as a function of η . It has the major
advantage of simplicity, yet refined solutions are reachable if the control allocation matrix also
gets optimized: this is the last method presented.

7.1 Development of a Three-Axes H∞ Model Reference Tracking


Scheme

This section is dedicated to describing a methodology for computing simultaneously longitudinal


and lateral-directional control laws gains. The RMT principle is introduced for longitudinal
synthesis, and refinements are explained.

7.1.1 Model Reference Tracking for Longitudinal Synthesis

7.1.1.1 Description of the Model Reference Tracking Scheme

The RMT scheme is an alternative formulation to the ASF for specifying desired closed-loop
handling qualities. Under it simplest form, it consists in minimizing the difference between a
reference model and the actual closed-loop, in the sense of the H∞ norm. For a comprehensive
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 151

development on RMT , readers may refer to [Sima, 1996].

The longitudinal RMT scheme is presented on figure 7.1. Pilot – or automatic pilot – inputs
are given in terms of commanded load factor nref
z . The desired closed-loop dynamics for tracking
this commanded load factor nref
z is defined through following transfer function:

nref
z ω0 2
= 2 (7.1)
n zc s + 2ξ0 ω0 s + ω0 2

That is a second order system whose frequency-domain and time-domain behavior are well-
known. Following values were chosen for the reference model:

• ω0 =1 rad.s−1

• ξ0 =0.7

Remark 7.1
The choice of adequate values for the closed-loop model is in itself an interesting question when
it comes to controlling a non-conventional airplane. Indeed traditionally these parameters are set
based on open-loop dynamics: this tends to limit the control amplitude in closed-loop. Yet when
the open-loop dynamics gets unusual, as it is in our case with two real poles, one stable and one
unstable, it is not straightforward to derive closed-loop objectives. In this report we chose to set
a constant objective whatever the flight point. Yet for further studies a more promising approach
could be to adapt closed-loop requirements from open-loop dynamics, even when the latter are
unusual, provided their frequency lie within a given frequency range.

The basic RMT problem is then:

min Tnzc →z1 (P, K) ∞


(7.2)
K
such that: K internally stabilizes P

with:
z1 = nref
z − nz (7.3)

and P (s) is the plant model defined on figure 7.1/ A major difference with previously introduced
ASF form is that with the RMT scheme, the ideal H∞ norm – attained if and only if the closed-
loop perfectly matches the reference model– is 0, and no more 1. As a consequence all H∞ optimal
norms should now be compared to 0 instead of 1. It also results, and this is sometimes pointed
out as a drawback of this technique, that optimal H∞ norms may no more be interpreted as a
percentage of overshooting a desired model, but can only be compared one to another.
152 Chapter 7. Co-Design of Control Surfaces and Laws

P (s)
ω02 nref + z1
z
s2 +2ξ0 ω0 s+ω02

nzc Knzc α
+
Aircraft
+ 1 Ki + δmequi Kalloc u 2
ωact
2
s2 +2ξact ωact s+ωact
Ua Longitudinal Ya Delay q
− s + Model nz
Actuators Model
+

Kq

Knz

Figure 7.1: Longitudinal model reference tracking scheme, with longitudinal gains structure and
error computed on the actual nz signal: z1 = nref
z − nz .

7.1.1.2 Structure for Longitudinal Gains

The control law structure used in this chapter is a C∗ law type, that is pitch stability is obtained
through a combination of pitch rate and load factor feedbacks. A feedforward gain is added, as
well as an integral term of the tracking error for precision purpose. For longitudinal synthesis there
are therefore four gains to optimize:

• Load factor feedback Knz

• Pitch rate feedback Kq

• Commanded load factor feedforward Knzc

• Integral gain Ki

This gains structure, typical of the Airbus fly-by-wire architecture, is described extensively in
[Favre, 1994]. It is also represented on figure 7.1. The output of the C∗ law is an equivalent
elevator order δmequi . This equivalent order is computed as follows:

Z
δmequi = Knzc nzc + Knz δnz + Kq q + Ki (nzc − nz ) (7.4)

7.1.1.3 Initial Longitudinal Synthesis

The RMT principle is now applied to the longitudinal case for the following example flight point:

• Mach 0.35
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 153

ref
Wpitch
Tnzc →nz

nref
z /nzc
nz /nzc
z1 /nzc

(a) Comparison of reference model, closed-loop and error (b) Comparison of longitudinal poles for refer-
frequency responses for a minimization of Tnzc →z1 ∞ . ence model and closed-loop, for a minimization of
Tnzc →z1 ∞ .

Figure 7.2: Bode diagram and poles representation on complex plane, for longitudinal reference
model and closed-loop resulting from minimization of Tnzc →z1 ∞ .

• m = 300T

• H = 3300f t

• η = 0.6

This is a low-speed, unstable flight point. It was chosen to demonstrate the method using a
value η = 0.6 for control surfaces span, first in order to show the ability of our method to cope
with different values of η parameter, and also to provide more significant deflections in temporal
responses than for η = 1 case.

For a fixed value of η the control allocation is a fixed matrix, which is chosen equal to
the pseudo-inverse of Bq = [. . . mδmi . . . ], similarly to what was considered in section 5.4. Con-
cerning actuators bandwidth, chapter 5 concluded in the feasibility of longitudinal stabilization
with moderate bandwidth: it was indeed shown that from a power consumption perspective it
is optimal to segregate actuators bandwidth for the different elevons, and to consider only one
“fast” actuator above 2.5Hz on the outer elevon. Yet in present chapter each control surface
area is varying while η changes, so bandwidth segregation does not make sense any more for
this study. As a consequence we now consider a unique bandwidth for all elevons. This unique
bandwidth is chosen slightly smaller than what was needed if only elevon 5 was actuated: a value
of ωact = 1.4Hz was then retained.

In order to provide a more accurate model, the actuators dynamics was also changed from a
first-order to a second-order, with a damping ξact = 0.7:

Ua ωact 2
= 2 (7.5)
u s + 2ξact ωact s + ωact 2
154 Chapter 7. Co-Design of Control Surfaces and Laws

1.5 Control Surfaces Response


Elevons Deflections (in ˚)

1
δmL1
0.5 δmL2
δmL3
0 δmL4
δmL5
δmR1
-0.5
δmR2
δmR3
-1 δmR4
δmR5
-1.5 δn
0 1 2 3 4 5 6 7 8 9 10
Time (s)
2 Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

0 ˙ L
δm 1
˙ L
δm 2
-1 ˙ L
δm 3
˙ L
δm 4
-2 ˙ L
δm 5
˙ R
δm 1
-3 ˙ R
δm 2
˙ R
δm 3
˙ R
δm
-4 ˙ R
4
δm 5
˙
δn
-5
0 1 2 3 4 5 6 7 8 9 10
Time (s)
1.2 Aircraft Response

0.8

0.6
δnz

0.4

0.2

0 nref
z
nz
-0.2
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 7.3: Temporal responses for a step input δnzc = 1g, with control law computed by
minimization of Tnzc →z1 ∞ .
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 155

A 100ms delay is also incorporated into the simulation model in order to account for delays in
the acquisition chain. This delay is approximated as a second-order Padé filter for the synthesis.

With control law structure introduced in section 7.1.1.2, the RMT problem becomes:

min Tnzc →z1 (P, Knzc , Ki , Knz , Kq ) ∞


(7.6)
Knzc ,Ki ,Knz ,Kq

such that: [Knzc , Ki , Knz , Kq ] internally stabilizes P

Similarly to previous chapters, this problem is solved using systune routine. Following optimal
value is obtained:
γ∞ = 0.174 (7.7)

Remark 7.2
In figure 7.3 and all following temporal responses, elevons deflections are plotted relatively to their
initial trimmed position. We remind that all control surfaces are used in an equal manner for
initial trim of the aircraft. These trimmed positions will yet be taken into account during the
optimization of section 7.3: deflections stops will be considered relatively to the initial trimmed
position of control surfaces.

Bode magnitude diagrams of the desired template nref z , the actual closed-loop nz and
the error z1 as a function of frequency are displayed on figure 7.2(a), together with a complex
plane representation of closed-loop poles on figure 7.2(b). On this figure we see that closed-loop
poles are located close to the reference model poles. A right-plane real zero comes from the
direct lift effect of elevons on load factor. The integral pole originating from the control law
structure is set to be on limit of stability by the optimizer: by default the optimizer ensures all
poles have a real part of at least 10−7 . This value is also visible on the “break” occurring at 10−7
of the Bode magnitude response of z1 . The near-zero position of the integrator has indeed the
detrimental effect of entailing a static error on the load factor δnz , visible on the temporal response
of figure 7.3. This simulation shows deflections, deflection rates and load factor in response to a
commanded step input δnzc = 1g – that is a commanded load factor nzc = 2g. Also visible on
the temporal simulation is the direct lift effect of elevons, translating from a control perspective
into a non-minimal phase zero: the initial effect of moving the elevons upward to achieve a higher
load factor is a decrease of total lift, implying an initial decrease of load factor. The elevons
kinematics is also typical from controlling an unstable aircraft: the final position of controls is
opposite to their initial displacement. Physically speaking we can say that in order to initiate the
pull-up elevons move upward –that is negative –, and then they have to move downward in order
to “stop” the movement, such latter move being different from that of a conventional airplane.

7.1.1.4 Improvement of the Method by Adding a Poles Constraint

In order to improve the solution, and to force the integral term to be active on a wider frequency
range, a constraint on closed-loop poles location is added. Following constraints are set:
156 Chapter 7. Co-Design of Control Surfaces and Laws

ref
Wpitch
Tnzc →nz

nref
z /nzc
nz /nzc
z1 /nzc

(a) Comparison of reference model, closed-loop and error (b) Comparison of longitudinal poles for reference model
frequency responses for a minimization of Tnzc →z1 ∞ , and closed-loop, for a minimization of Tnzc →z1 ∞ , with
with a constraint set on poles location. a constraint set on poles location.

Figure 7.4: Bode diagram and poles representation on complex plane, for longitudinal reference
model and closed-loop resulting from minimization of Tnzc →z1 ∞ with a constraint set on poles
location.

• MinDecay=0.2 rad.s−1

• MinDamping=0.5

These are constraints that ensure following mathematical inequalities:

∀p ∈ C, p pole of P (s) : (7.8)


Re(p) ≤ −MinDecay (7.9)
Re(p) ≤ −MinDamping.|p| (7.10)

The constraint set on minimal poles damping may seem redundant with specifications of the
reference model. Yet we chose to add these constraint in order to limit the domain of solutions
for the optimizer, and to ensure a good damping of the closed-loop responses.

A visualization of these constraints is provided on figure 7.4(b). The constrained optimization


problem then becomes:
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 157

1.5 Control Surfaces Response


Elevons Deflections (in ˚)

1
δmL1
0.5 δmL2
δmL3
0 δmL4
δmL5
δmR1
-0.5
δmR2
δmR3
-1 δmR4
δmR5
-1.5 δn
0 1 2 3 4 5 6 7 8 9 10
Time (s)
2 Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

0 ˙ L
δm 1
˙ L
δm 2
-1 ˙ L
δm 3
˙ L
δm 4
-2 ˙ L
δm 5
˙ R
δm 1
-3 ˙ R
δm 2
˙ R
δm 3
˙ R
δm
-4 ˙ R
4
δm 5
˙
δn
-5
0 1 2 3 4 5 6 7 8 9 10
Time (s)
1.2 Aircraft Response

0.8

0.6
δnz

0.4

0.2

0 nref
z
nz
-0.2
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 7.5: Temporal responses for a step input δnzc = 1g, with control law computed by
minimization of Tnzc →z1 ∞ and adding a constraint on poles location.
158 Chapter 7. Co-Design of Control Surfaces and Laws

min Tnzc →z1 (P, Knzc , Ki , Knz , Kq ) ∞


(7.11)
Knzc ,Ki ,Knz ,Kq

such that:
∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
[Knzc , Ki , Knz , Kq ] internally stabilizes P

The optimal value obtained is:

γ∞ = 0.187 (7.12)
gBest = 0.9934 (7.13)

Remark 7.3
gBest represents the maximum of all normalized constraints of the optimization problem. If
constraints are set on H∞ norms, then gBest is the normalized H∞ norm at the optimum. For
constraints on poles, it represents the distance to the constraint. A value of gBest below 1 then
means all constraints are fulfilled.

Comparing γ∞ = 0.187 to equation (7.7) shows a slight increase of the optimal value when
adding constraints, compared to the pure minimization problem. However on figures 7.4(a)
and 7.4(b) we see that the integral term is now set at 0.2 rad.s−1 . This has the expected
advantage of providing a quasi-zero steady-state error, without affecting the template matching
at higher frequencies. This can also be observed on temporal simulation of figure 7.5: transient
dynamics is similar, however the steady-state error on load factor is now canceled.

7.1.1.5 Further Improvement of the Method by Changing Tracking Signal

We now have a satisfactory formulation of the model reference tracking problem for longitudinal
synthesis. However a last improvement may be added to the formulation: indeed the optimal
value γ∞ = 0.187 appears too high, especially when compared to optimal values on lateral and
directional axes that will be developed in section 7.1.2. In fact, it comes out that this high value
originates from the non-minimal phase zero of the plant, on which the control law has no degree
of freedom to act. As a consequence this zero strongly penalizes the optimal value.

To improve this, and in order to prepare for the three-axes H∞ synthesis, a new channel
is substituted to the load factor for the computation of the reference tracking error: instead of
minimizing the error between the reference model nref
z and the actual load factor nz , the error
is now computed using an approximated load factor ñz which does not include the non-minimum
phase zero. Instead δñz is computed from relative angle-of-attack δα:
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 159

ω02 nref + z̃1


z
s2 +2ξ0 ω0 s+ω02

ñz

nzc Knzc α Ve
+ g α
z
Aircraft
+ 1 Ki + δmequi Kalloc u 2
ωact
2
s2 +2ξact ωact s+ωact
Ua Longitudinal Ya Delay q
− s + Model nz
Actuators Model
+

Kq

Knz

Figure 7.6: Longitudinal model reference tracking scheme, with longitudinal gains structure and
error computed on approximated ñz signal: z̃1 = nref
z − ñz .

DCgain(Tnzc →nz )
δñz = δα (7.14)
DCgain(Tnzc →α )
Ve
= zα δα (7.15)
g

Once again the rationale is to consider that while the control law has no effect on the elevons
direct lift, then the best from the optimization point of view is just to remove this zero from
the synthesis. Of course in the actual simulation the load factor remains the “real” one, not the
approximated one.

Remark 7.4
Asserting that the control law has no effect on direct lift phenomenon is strictly speaking a bit
short. Some strategies could indeed be investigated, which could provide direct lift control taking
advantage of the elevons redundancy. From a control perspective this would mean adding a lift
target into the control allocation problem, and finding a control allocation strategy that fulfills
both moments and lift requirements. From a physical perspective it implies playing with the
difference of x−wise lever arm of all control surfaces. For instance we could imagine dynamically
counter-acting the direct down-lift effect of a control surface by moving down a less pitch-efficient
control surface. Of course this method would be quite control-consuming and should therefore be
limited to small orders, yet it should certainly be investigated for some critical flight phases such
as landing.

The augmented RMT scheme with inclusion of approximated load factor ñz and resulting
tracking error z̃1 = nref
z − ñz is displayed on figure 7.6. The constrained optimization problem is
then:
160 Chapter 7. Co-Design of Control Surfaces and Laws

ref
Wpitch
Tnzc →nz

nref
z /nzc
nz /nzc
z1 /nzc

(a) Comparison of reference model, closed-loop and error (b) Comparison of longitudinal poles for reference model
frequency responses for a minimization of Tnzc →z̃1 ∞ , and closed-loop, for a minimization of Tnzc →z̃1 ∞ , with
with a constraint set on poles location. a constraint set on poles location.

Figure 7.7: Bode diagram and poles representation on complex plane, for longitudinal reference
model and closed-loop resulting from minimization of Tnzc →z̃1 ∞ with a constraint set on poles
location.

min Tnzc →z̃1 (P, Knzc , Ki , Knz , Kq ) ∞


(7.16)
Knzc ,Ki ,Knz ,Kq

such that:
∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
[Knzc , Ki , Knz , Kq ] internally stabilizes P

The optimal value obtained is:

γ∞ = 0.0853 (7.17)
gBest = 0.9987 (7.18)

A reduction factor of more than 2 is achieved with this formulation compared to previously
found value γ∞ = 0.187. As it can be observed on frequency responses of figure 7.7 and tem-
poral simulation of figure 7.8, this enhancement is not obtained at the expense of a performance
degradation of the control law, on the contrary. Appart from the initial down-lift effect, the nref
z
model is closely followed.

As a conclusion for this part, a step-by-step approach was developed in order to construct
a simple yet effective model reference tracking scheme for longitudinal synthesis. This scheme
will now be a basis for the three-axes RMT scheme developed hereafter, and ultimately for the
three-axes codesign.
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 161

1.5 Control Surfaces Response


Elevons Deflections (in ˚)

1
δmL1
0.5 δmL2
δmL3
0 δmL4
δmL5
δmR1
-0.5
δmR2
δmR3
-1 δmR4
δmR5
-1.5 δn
0 1 2 3 4 5 6 7 8 9 10
Time (s)
2 Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

0 ˙ L
δm 1
˙ L
δm 2
-1 ˙ L
δm 3
˙ L
δm 4
-2 ˙ L
δm 5
˙ R
δm 1
-3 ˙ R
δm 2
˙ R
δm 3
˙ R
δm
-4 ˙ R
4
δm 5
˙
δn
-5
0 1 2 3 4 5 6 7 8 9 10
Time (s)
1.2 Aircraft Response

0.8

0.6
δnz

0.4

0.2

0 nref
z
nz
-0.2
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 7.8: Temporal responses for a step input δnzc = 1g, with control law computed by
minimization of Tnzc →z̃1 ∞ and adding a constraint on poles location.
162 Chapter 7. Co-Design of Control Surfaces and Laws

ω02 nref
z + z̃1
s2 +2ξ0 ω0 s+ω02 −
Reference pitch dynamics
1 φref + z2
(1+τrp s)(1+τsp s)

Reference roll dynamics
2
ωdr βref + z3
2
s2 +2ξdr ωdr s+ωdr

Reference yaw dynamics

η η α Ve
g zα
nz
q
LFR
δmequi LFR Aircraft β
nzc C ∗ Law Pseudo-inverse 2
ωact u
δmi i = 1..10 Representation
Allocation δn s2 +2ξ
act ωact s+ωact
2 Delay p
φc δlequi Actuators model r
βc Y ∗ Law δnequi ew q 1+ V

3Lz
s wz φ
σz 2L z e
πVe (1+ Lz s)2
Ve

Turbulence model

Figure 7.9: Three-axes RMT scheme. Tunable blocks are displayed in orange.

7.1.2 Reference Model Tracking for Three-Axes Synthesis

This section is now dedicated to describing the general three-axes RMT scheme which is af-
terwards used for the codesign optimization. First the problem is formulated, then the general
structure of the lateral-directional control law is presented, and the performance of the control
law is analyzed.

7.1.2.1 Three-Axes Formulation of the Reference Model Tracking Scheme

The three-axes RMT scheme we used is very similar to the previously developed longitudinal
case. The main change is that now the transfer to be minimized is no more a SISO transfer but a
MIMO one. More precisely pilot – or automatic pilot – orders are given in terms of commanded
load factor nzc , bank angle φc 1 and sideslip angle βc . For each axis a reference dynamics is
specified:

• Concerning the roll axis, we seek a combination of two first-order responses:

φref 1
= (7.19)
φc (1 + τRP s)(1 + τSP s)
1
Here there is a slight discrepancy between our implementation and the Airbus fly-by-wire philosophy: roll inputs
from a stick pilot perspective a usually roll rate orders, while objectives in terms of commanded bank angle are
automatic pilot objectives. Changing this structure would be a subject for future work. As a consequence, some
conclusions we draw from our study could possibly be affected by this change.
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 163

• Concerning the yaw axis, we seek to damp the dutch roll mode as follows:

βref ωdr 2
= 2 (7.20)
βc s + 2ξdr ωdr s + ωdr 2

Values for reference specifications are provided on table B.1. These values were derived from
open-loop values on a low-speed case. Similarly to what was proposed for longitudinal synthesis,
these specifications were kept constant for all flight points in this study. As a result for flight
points where there is a significant gap between open and closed-loop, overly important control
amplitude may result. A promising enhancement for future studies would be to apply the same
approach while adapting reference models to the flight points.

Parameter τrp τsp ωdr ξdr


Value 1.6 rad.s−1 1.9 rad.s−1 0.4 rad.s−1 0.7

Table 7.1: Values for lateral and directional reference models.

Instead of specifying a desired performance axis by axis, the objective is set on the 3 × 3
transfer from pilot inputs to tracking errors. The unique minimization objective of the three-axes
RMT scheme is written as follows:

min T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P, K) (7.21)


K ∞

Here model-reference tracking is written as minimizing the H∞ norm of a three-input three-


output transfer function between pilot inputs (nzc , φc , βc ) and outputs (z̃1 , z2 , z3 ) the differences
between reference dynamics outputs and actual closed-loop signals (ñz , φ, β). For a proper def-
inition of all signals please refer to figure 7.9. In our case multi-channel transfer has several
advantages over multiple SISO transfers. First off-diagonal terms are implicitly set to zero.
Hence the resulting control law will totally decouple all three axes, namely longitudinal from lat-
eral / directional which is obvious given the aircraft dynamics around a longitudinal equilibrium,
but also lateral from directional – turns are performed with zero sideslip – and directional from
lateral – “pedal” inputs imply no bank –. Couplings between all axes are explicitly taken into
account by the control law, which a SISO approach would not. Then this formulation is virtually
independent from the open-loop model and structure of control laws. Finally it allows for an
elegant formulation of the optimization problem of section section 7.3: a single constraint on the
H∞ norm of this multi-channel transfer ensures an adequate closed-loop behavior on all three
axes. An equivalent formulation with SISO transfers would require nine constraints.

However a practical requirement to ensure the success of such a formulation is that optimal
H∞ values of each channel have similar order of magnitude. Otherwise the min-max optimizer
will focus on minimizing the largest signal, at the expense of increasing values on other channels.
Such a normalization may be obtained by using weighting matrices for each channel, yet finding a
unique weight valid for all flight points is a tricky task. More elegantly, channels should if possible
be chosen to have directly the same order of magnitude in order to be comparable: this was the
point of introducing the approximated load factor ñz in section 7.1.1.5.
164 Chapter 7. Co-Design of Control Surfaces and Laws

7.1.2.2 Structure for Lateral Gains

Gains structure for longitudinal gains is retained from section 7.1.1.2. Structure for lateral gains
is again typical from the Airbus fly-by-wire philosophy, which is extensively described in [Favre,
1994]: so-called Y ∗ structure features lateral / directional state feedback, namely sideslip β ,
yaw rate r , bank angle φ and roll rate p . An integrator is added to keep zero steady-state
sideslip, as well as a bank angle and sideslip orders direct feedthrough gains. Outputs of this
law are equivalent aileron and rudder order δlequi and δnequi respectively. These outputs are
computed as follows:

Z
δlequi = Kφc φc + Kpβ β + Kpp p + Kpr r + Kpφ φ + Kipβ (βc − β) (7.22)
Z
δnequi = βc βc + Krβ β + Krp p + Krr r + Krφ φ + Kirβ (βc − β) (7.23)

Finally the control problem features 16 variables, gathered in the vector K :

K =(Knzc , Knz , Kq , Ki , . . . (7.24)


Kpβ , Kpp , Kpr , Kpφ , Kipβ , Kφc , . . .
Kβc , Krβ , Krp , Krr , Krφ , Kirβ )

7.1.2.3 LFR Parametrization of the Control Allocation Matrix by Control Surfaces Span
Parameter

We remind that control allocation problem is that of finding a deflections vector u satisfying:
   
Cmδm1 ... Cmδm10 Cmδn Cmδmequi δmequi
 Clδm1 ... Clδm10 Clδn  u =  Clδlequi δlequi  (7.25)
   

Cnδm1 ... Cnδn10 Cnδn Cnδnequi δnequi


| {z }
B1 (η)

where B1 (η) is the matrix of elevons gradients in pitch, roll and yaw respectively, which all
depend from parameter η . [Cmδmequi Clδlequi Cnδnequi ]T is a vector of equivalent gradients
as seen by the control law. These values may be set arbitrarily without loss of generality, we
chose equivalent values of 1 on all axes. Then a classical solution of equation (B.12) is the
Moore-Penrose pseudo-inverse:
 
δmequi
u = Kalloc (η)  δlequi  (7.26)
 

δnequi
with Kalloc (η) = B1T (B1 B1T )−1 (7.27)
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 165

As moments gradients depend on η parameter, this dependency should be taken into account
in the allocation matrix Kalloc (η). For that purpose an LFR approximation of the allocation
matrix for varying parameter η was computed from sampled pseudo-inverses for discretized
values of η , with a similar methodology as presented in chapter 6. Resulting allocation gains
for different elevons as a function of η are shown on figure 7.10. In a general way absolute
norm of gains increase while η decreases. This may seem counter-intuitive, however this only
means that when control surfaces sizes decrease their deflection should be increased in order
to keep a constant equivalent pitch and roll efficiency. Similarly control allocation gains on
elevon 1 increase in absolute value whereas this control surface remains fixed. This is due to
its higher relative efficiency when other elevons size decreases. Resulting three-axes model with
parameterized control allocation for a fixed value of η is shown on figure 7.9.

Parameterizing the control allocation matrix with the parameter η and computing a con-
tinuous approximation using the LFR framework is a simple yet powerful idea we developed:
it allows taking into account the variation of geometrical parameters in an explicit way into the
control laws. This idea will enable two alternatives to the codesign problem in section 7.3.

0 -0.1 2.5
2
2
-0.2
1.5
-0.2 1.5
alloc
Kpitch
alloc

Kroll

-0.3
1 1
-0.4 -0.4
0.5
0.5
-0.5
-0.6 1 0 1
1 0.5 1 2 0.5
2 3 η 3
4 5 0 Elevon 4 5 0 η
Elevon

pitch roll
(a) Variation of pitch allocation gains Kalloc for vary- (b) Variation of roll allocation gains Kalloc for varying
ing η . η.

Figure 7.10: LFR polynomial approximations of control allocation matrix Kalloc .

7.1.2.4 Performance Evaluation on an Example Flight Point

After having described the different features of three-axes RMT control, the result of a synthesis
on the same example flight point as developed for the longitudinal case is examined here. The
general H∞ problem of equation (7.21), taking into account the gains structure of equation (7.24)
and poles constraint introduced in equation (7.11), reads:

min T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η), K) (7.28)


K ∞
such that:
∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η init )
166 Chapter 7. Co-Design of Control Surfaces and Laws

Step Response
From: nzc From: φc From: βc
To: nz

Reference Model
Closed-loop
Amplitude
To: φ
To: β

Time (seconds)

Figure 7.11: Comparison of reference models and temporal responses on nz , φ and β to a


unit step input in nzc , φc and βc respectively.

After optimization, following optimal values are obtained:

γ∞ = 0.107 (7.29)
gBest = 0.9997

The γ∞ value should be compared to that obtained for the pure longitudinal minimization,
in equation (7.17). A 25% increase is observed compared to the longitudinal case: this indicates
that the longitudinal channel is not the limiting channel from a H∞ optimal perspective, at least
on this flight point. This observation is corroborated by the analysis of Bode magnitude diagrams
channel by channel of figure 7.12. Interestingly the peak amplitude is on the bank angle channel,
moreover it is located in low frequency. This possibly indicates a static error on the commanded
bank angle.

Examining responses of each channel to an unitary step input on figure 7.11 confirms the
presence of a slight static error on φ . This is due to the lack of any integral term of the error
in φ in the gains structure introduced in section 7.1.2.2. Adding an integral term would have
improved the situation, however we restricted to stick with the structure coming from industrial
requirements. As already suggested, an alternative could also be to work with commanded roll
rate instead of bank angle. Apart from the static error in φ , all reference models are precisely
tracked, and decoupling requirements are perfectly matched.
7.1. Development of a Three-Axes H∞ Model Reference Tracking Scheme 167

nref
z /nzc φref /φc
nz /nzc φ/φc
z1 /nzc z2 /φc

(a) [Comparison of longitudinal reference model, closed- (b) [Comparison of lateral reference model, closed-loop
loop and error frequency responses for a minimization and error frequency responses for a minimization of
of T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } ∞ , with a constraint set on T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } ∞ , with a constraint set on
poles location. poles location.

βref /βc
β/βc
z3 /βc

(c) [Comparison of directional reference model, closed-


loop and error frequency responses for a minimization
of T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } ∞ , with a constraint set on
poles location.

Figure 7.12: Bode diagrams for three-axes reference models and closed-loop resulting from mini-
mization of T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } and with a constraint on closed-loop poles location.

A last simulation is performed in order to examine the performance of the control law: at
t = 0s, a pull-up δnzc = 1g is commanded, then at t = 5s a bank angle φc = 30˚ is initiated, and
finally at t = 12s a pedal input βc = 5˚ is commanded. The simulation is visible on figure 7.13. In
line with previous step response, all pilot inputs are well tracked, with an adequate decoupling. A
2˚ static error is visible on φ . As expected, the control allocation module deflects symmetrically
the elevons for a longitudinal maneuver, and anti-symmetrically for a lateral maneuver. More
effective elevons on each axis are more deflected: namely elevon 1 in longitudinal and elevon 2
168 Chapter 7. Co-Design of Control Surfaces and Laws

Control Surfaces Response


Elevons Deflections (in ˚)

δmL1
δmL2
δmL3
δmL4
δmL5
δmR1
δmR2
δmR3
δmR4
δmR5
δn
Time (s)
Control Surfaces Deflection Rates

˙ L
δm
Eflection Rates ((in ˚/s)

1
˙ L
δm 2
˙ L
δm 3
˙ L
δm 4
˙ L
δm 5
˙ R
δm 1
˙ R
δm 2
˙ R
δm 3
˙ R
δm 4
˙ R
δm 5
˙
δn

Time (s)
Aircraft Response

q
q,β,p,r,φ

β
p
δnz

r
φ
nz

Time (s)

Figure 7.13: Temporal responses to inputs of δnzc = 1g at t = 0s, φc = 30˚ at t = 5s, and
βc = 5˚ at t = 12s, with gains computed by a minimization of T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } under

constraint of closed-loop poles location.
7.2. Translation of Handling Qualities Dynamic Constraints into H∞ Constraints 169

in lateral. Deflections amplitude and deflection rates are relatively important: let us just remind
that this simulation was performed with only half of outer elevon span available (η = 0.5). A bit
surprisingly, higher deflection rates are obtained for the turn maneuver rather than for the pull-up
maneuver. As we will see afterwards, what is visible on this example is in fact a generality for the
configuration.

Finally we notice an unexpectedly high deflection and rate for the rudder (which is in fact,
as discussed in section 2.1.5, the combination of two separate rudders gathered into a single
equivalent effector with twice the effectiveness of the two separate rudders), especially at the
beginning of the turn maneuver in order to keep turn coordination. This may first indicate that
the turn coordination law is too efficiently tuned: perhaps a perfectly zero sideslip is not required
during the whole temporal maneuver, and the rudder deflection could be lowered without much
effect on the aircraft handling qualities. However this could also indicate a possible under-sizing of
the fin and rudder: while this is out of scope of our study, it would deserve a closer look in further
feasibility studies. Turn coordination is indeed generally considered as a good way to estimate the
directional capabilities of an aircraft.

As a conclusion for this part, we have set up an elegant formulation for a simultaneous
optimization of all control laws gains for longitudinal-lateral control. The main advantage of this
formulation is its conciseness: only one objective and one constraint are sufficient in order to
track desired closed-loop models. This conciseness will be used in the next part: putting the
H∞ objective as a constraint will ensure adequate handling qualities: in other words it is now
possible to express the three-axes closed-loop handling qualities problem into a single constraint
satisfaction problem 2 . While better performance on each individual channel may be obtained by
separately optimizing each channel, this is not the aim of our work: instead of seeking the best
achievable control law, we seek a compromise between good tracking and a light formulation, with
the codesign perspective in mind.

7.2 Translation of Handling Qualities Dynamic Constraints into


H∞ Constraints

A major step towards the codesign of parameter η is expressing handling qualities constraints
under a form understandable by the optimizer. Here all handling qualities constraints are expressed
as maximum H∞ norms of transfers for given inputs: it means all quantities are Root Mean-
Square (RMS) (see the definition of the H∞ norm as a maximum of the quotient of two RMS
in equation (3.22)). This approach is based on the work published in [Niewoehner and Kaminer,
1996; Kaminer et al., 1997]. How this can be interpreted, and what are the chosen handling
qualities constraints is explained now.

Three different types of constraints are chosen in order to size the control surfaces size η :

• Maximum deflections and deflection rates in response to a pilot pull-up. This is ensured
2
2 constraints if including the poles constraint.
170 Chapter 7. Co-Design of Control Surfaces and Laws

through constraint on following H∞ norms for each control surface:

1
Tn →δmi ∆nzc max ≤1 (7.30)
∆δmi max zc ∞
1
Tnz →δm˙ i ∆nzc max ≤1 (7.31)
˙
δmi max c

with ∆δmi max = δmi max − δmie .

• Maximum deflections and deflection rates in response to a severe continuous longitudinal


turbulence, whose model was introduced in section 2.5. This is ensured through constraint
on following H∞ norms for each control surface:

2
Te →δmi ≤1 (7.32)
∆δmi max w ∞
2
T ˙ ≤1 (7.33)
δmi max ew →δmi
˙ ∞

• Maximum deflections and deflection rates in response to a pilot bank angle order. This is
ensured through constraint on following H∞ norms for each control surface:

1
Tφ →δmi φc max ≤1 (7.34)
∆δmi max c ∞
1
T φ max ≤1 (7.35)
˙ i max φc →δm˙ i c
δm ∞

Chosen values for the sizing are:

• δmi max = 25˚. A 5˚margin versus deflection stops at 30 ˚was chosen, to avoid entering the
non-linear part of control authority. Please also note that for the maximal deflection allowed,
the deflection used for trim δmie is taken into account into the computation of ∆δmi max :
the more deflections are needed in order to balance the aircraft, the less authority remains
available for control.
˙ i max = 60˚/s.
• δm

• ∆nzc max = 1.5g.

• φc max = 45˚.

As already pointed out these constraints are frequency-domain requirements: no guarantee


can be obtained in time-domain simulations. However this approach is largely spread in literature
(see for instance [Niewoehner and Kaminer, 1996; Yang and Lum, 2003; Silvestre et al., 1998]),
and we believe it still delivers useful information for sizing at future projects level. What these
constraints mean is now developed in two codesign examples.

First example is finding the minimum control surfaces size η such that a φc = 30˚ bank angle
order does not imply more than δmi max = ±10˚ deflection. The problem formally reads:
7.2. Translation of Handling Qualities Dynamic Constraints into H∞ Constraints 171

min η (7.36)
K,η

such that:
1
Tφ →δmi 30˚ ≤ 1∀i = 1 . . . 5 (7.37)
10˚ c ∞
T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η), K) ≤ 1.05 ∗ γ∞ (7.38)

∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η)

In equation (7.38), γ∞ is computed from the optimization presented in section 7.1.2; here
γ∞ = 0.107. The 1.05 factor is a relaxation factor that ensures the optimizer does not get
stuck into a local minimum. The codesign problem has here 7 constraints: the 5 first constraints
are associated with constraints on each control surface deflection, while the latter two are the
H∞ constraint and poles constraint respectively, ensuring a correct closed-loop behavior. After
optimization an optimal value η opt = 0.4165 is found. Constraints at the optimum are represented
on figure 7.14. We see that the second elevon has its constraint saturated. The order of elevons
saturations is given by the control allocation, which is chosen parameterized by η as described
in previous section. The H∞ constraint on RMT is also saturated.

kTφc →δm1 . 30˚k 30˚


kTφc →δm2 . 10 kTφc →δm3 . 30˚k
10˚ ∞ ˚k∞ 10˚ ∞

kTφc →δm4 . 30˚k 30˚


kTφc →δm5 . 10 1
10˚ ∞ ˚k∞ kT{nzc ,φc ,βc }→{z1 ,z2 ,z3 } . 1.03∗γ∞
k∞

Figure 7.14: H∞ constraints from φc to δmi normalized.


172 Chapter 7. Co-Design of Control Surfaces and Laws

We also notice that the peak amplitude is achieved for a frequency ω = 0.79 rad.s−1 . A
temporal simulation of control surfaces deflection in response to an input φc = 30˚sin(ωt) with
ω = 0.79 rad.s−1 and η opt = 0.4165 is displayed on figure 7.15. The 10˚ limit on second elevon
deflection is achieved in steady state oscillatory regime.

Control Surfaces Response

δmL
1
δmL
2
Elevons Deflections (in ˚)

δmL
3
δmL
4
δmL
5
δmR
1
δmR
2
δmR
3
δmR
4
δmR
5

Time (s)

Figure 7.15: Control surfaces deflections to a sinusoidal input of φc at frequency 0.79 rad.s−1 ,
for η opt = 0.4165 computed so that ∆δm1i max Tφc →u φmax ≤ 1, with ∆δmi max = 10˚ and

φmax = 30˚.

Then deflections amplitude in response to a step input φc = 30˚ is plotted on figure 7.16.
Maximal deflection of elevon 2 is a bit below 10˚. Once again nothing can be said whether
maximal amplitude is below or above prescribed limit in the general case. However maximal
amplitude is in the expected order of magnitude. This type of frequency domain specifications
has moreover the main advantage of being very efficient from a computational time perspective,
compared to simulation-based optimization. Therefore it is found suitable from a future projects
perspective.

A last example is performed on sizing based on response to longitudinal turbulence. This


time only maximal amplitude of 5˚in response to turbulence is considered, and the optimization
example writes:
7.2. Translation of Handling Qualities Dynamic Constraints into H∞ Constraints 173

Control Surfaces Response


Elevons Deflections (in ˚)

δmL1
δmL2
δmL3
δmL4
δmL5
δmR1
δmR2
δmR3
δmR4
δmR5
δn

Time (s)

Figure 7.16: Control surfaces deflections to a step input of φc at t = 5s, for η opt = 0.4165
computed so that ∆δm1i max Tφc →u φmax ≤ 1, with ∆δmi max = 10˚ and φmax = 30˚.

Control Surfaces Response


δm1
δm2
Elevons Deflections (in ˚)

δm3
δm4
δm5

Time (s)

Figure 7.17: Temporal response of control surfaces to severe longitudinal turbulence, with η opt =
0.3966 computed so that δmi2max Tew →u ≤ 1, with δmi max = 5˚.

174 Chapter 7. Co-Design of Control Surfaces and Laws

min η (7.39)
K,η

such that:
2
Te →δmi ≤ 1∀i = 1 . . . 5 (7.40)
5˚ w ∞
T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η), K) ≤ 1.05 ∗ γ∞ (7.41)

∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η)

This time an optimal value η opt = 0.3966 is found, with elevon 1 being limiting. A temporal
response to a severe longitudinal turbulence input is presented on figure 7.17. As introduced in
section 2.5, the Dryden turbulence model is a white-noise input passed through a low-pass filter,
with a variance σz = 5m.s−1 . As the variance value is exceeded in temporal response of wz , so
would be if a variance of 5˚ were specified on deflections. In order to limit excursions beyond 5˚,
we then chose to set the limit of 5˚ on the second quartile of the spectral distribution of δmi ,
that is 2σ = 5˚: hence the 2 factor in equation (7.32). As a consequence, on figure 7.17 we see
that the limit value of ±5˚ is overall held, with a slight overshoot at t = 30s.

7.3 Evaluation of Different Methods for Control Surfaces Codesign

Three different codesign methods are now explained, and applied on the full flight envelope. These
three methods range from the simplest to the most advanced: namely “decoupled” optimization,
optimization with parameterization of the allocation, and optimization with tunable allocation are
investigated.

7.3.1 Decoupled Optimization with LFR Parametrization of the Allocation

Decoupled optimization is explained, then results on the example flight point are analyzed, and
finally this method is applied on the entire flight envelope in order to find the sizing case for the
control surfaces span.

7.3.1.1 Problem Formulation

The basic idea for decoupled optimization is that thanks to the LFR parameterization of the
control allocation module described in section 7.1.2.3, at first order the C ∗ /Y ∗ control laws gains
are decoupled from the control surfaces span parameter η . So first control law gains are computed
7.3. Evaluation of Different Methods for Control Surfaces Codesign 175

ω02 nref
z + z̃1
s2 +2ξ0 ω0 s+ω02 −
Reference pitch dynamics
1 φref + z2
(1+s/τrp )(1+s/τsp )

Reference roll dynamics
2
ωdr βref + z3
2
s2 +2ξdr ωdr s+ωdr

Reference yaw dynamics

η η α Ve
g zα
nz
q
LFR
δmequi LFR Aircraft β
nzc C ∗ Law Pseudo-inverse 2
ωact u
δmi i = 1..10 Representation
Allocation δn s2 +2ξ
act ωact s+ωact
2 Delay p
φc δlequi Actuators model r
βc Y ∗ Law δnequi ew q 1+ V

3Lz
s wz φ
σz 2L z e
πVe (1+ Lz s)2
Ve

Turbulence model

Figure 7.18: Scheme for decoupled optimization. Tunable blocks are displayed in orange.

for a fixed η value using the procedure described in section 7.1.2, and then with these gains
fixed η is varied in order to fulfill handling qualities requirements described in previous section.
Scheme for this procedure is provided on figure 7.18. Mathematically the procedure reads:

Step 1:

(γ∞ , K opt ) ← min T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η init ), K) (7.42)
K ∞
such that:
∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η init )
176 Chapter 7. Co-Design of Control Surfaces and Laws

Step 2:

η opt ← min η (7.43)


η

such that:
1
Tn →δmi ∆nzc max (P (η), K opt ) ≤1
∆δmi max zc ∞
1 max
T ˙ ∆nzc (P (η), K opt ) ≤1
δmi max nzc →δmi
˙ ∞
2
Te →δmi (P (η), K opt ) ≤1
∆δmi max w ∞
2
Tew →δm˙ i (P (η), K opt ) ≤1
˙
δmi max

1
Tφ →δmi φc max (P (η), K opt ) ≤1
∆δmi max c ∞
1 max
T ˙ φc (P (η), K opt ) ≤1
δmi max φc →δmi
˙ ∞

While the first step involves one objective, 16 variables (the 16 control law gains) and one
constraint, the second step involves only one variable, one objective, and 30 constraints.

7.3.1.2 Results on a Single Flight Point

On the flight point M.35, H = 3300f t, m = 300T , following results are obtained:

γ∞ = 0.0844 for η init = 1 (7.44)


opt
η = 0.7415 (7.45)

Constraints values at the optimum are gathered in table 7.2.

Elevon number 1 2 3 4 5
1 max (P (η opt , K opt ))
∆δmi max Tnzc →δmi ∆nzc ∞
0.0911 0.0994 0.1058 0.0914 0.0861
1 max (P (η opt , K opt ))
˙
δmi max T nzc →δmi ˙ ∆n z c 0.0551 0.0601 0.0640 0.0552 0.0521

2 opt , K opt ))
T
∆δmi max ew →δmi (P (η 0.1874 0.2047 0.2178 0.1881 0.1772

2 opt , K opt ))
T
˙ i max ew →δmi
δm ˙ (P (η 0.1851 0.2021 0.2150 0.1857 0.1750

1 max (P (η opt , K opt ))
∆δmi max T φ c →δm i
φ c 0.1191 0.4225 0.5057 0.2999 0.1390

1 max (P (η opt , K opt ))
T
˙ i max φc →δm
δm ˙i cφ 0.2357 0.8356 1.0000 0.5931 0.2748

Table 7.2: Constraints values after decoupled optimization.

These constraints values after decoupled optimization are also displayed in figure 7.19. On this
figure we see that a single constraint over 30 is saturated: it corresponds to maximum deflection
7.3. Evaluation of Different Methods for Control Surfaces Codesign 177

rate in response to a commanded bank angle for control surface 3. Once again the order of control
surfaces saturation is given by the allocation. It is also noteworthy to remark that roll constraints
exceed by an order of magnitude longitudinal constraints. Whereas sizing for roll criteria is often
considered as secondary on BWB given the expected large roll authority provided by full-span
trailing edge control, surprisingly on this flight point roll sizing is predominant over pitch sizing.
This will be confirmed later on a majority of flight points.

0.9

0.8

0.7

0.6
H∞ norms of constraints

0.5

0.4

0.3

0.2

0.1

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
˙i
δm max T nzc →δmi˙ ∆n z c

2 opt , K opt ))
T
∆δmi max ew →δmi (P (η

2 opt , K opt ))
T
˙ i max ew →δm
δm ˙i (P (η

1 max (P (η opt , K opt ))
∆δmi max T φ c →δm i
φ c

1 max (P (η opt , K opt ))
T
˙ i max φc →δm
δm
φ
˙i c

Figure 7.19: Bar of constraints values after decoupled optimization, η opt = 0.7415.
178 Chapter 7. Co-Design of Control Surfaces and Laws

Comparison of temporal responses with η init = 1 and η opt = 0.7415 is provided on figure 7.20.
Interestingly temporal responses of the aircraft states are very similar for both layouts: let us remind
that here gains were computed once and for all for η init = 1, and the simulation for η = 0.7415
was performed changing only η value with the same gains. Only control surfaces amplitude and
deflection rates increases, to provide a similar outputs response. This validates the principle for
decoupled optimization: from the control law perspective, parametrization of the allocation makes
the system nearly independent from η from an input-output perspective 3 .

7.3.1.3 Decoupled Optimization on the Entire Flight Envelope

This decoupled optimization method is then run on the entire flight envelope, in order to determine
the sizing cases. This flight envelope features 156 flight points, with varying mass, Mach and
altitude at backward CG , as described in section 2.3. On figure 7.21 the amount of times
each constraint is limiting is plotted. A constraint is considered limiting if after optimization it
exceeds 0.99. The sum of all exceeding constraints may exceed the number of flight points: it
simply means several constraints may be limiting together for a single flight point. Clearly the
vast majority of flight points are sized by the roll constraint, more precisely by the deflection rates
constraint for a 30˚ bank angle input. This constraint is active most of the time on elevon 2 and
3: these are the most solicited by the allocation, depending on the η value (when η decreases,
both elevons 2 and 3 have increasing y−wise lever arm and decreasing areas. For small values of
η elevon 3 becomes more roll-effective than elevon 2, as it can be seen on figure 7.10(b)). A sum
of all these bars gives 190: it means 21% only of the flight points are sized by only one criterion,
for only one control surface. A possible improvement towards more optimality would be to seek
increasing this figure. This is achieved in next sections.

Then figure 7.22 shows η opt values after decoupled optimization on all flight points, and for
each flight point which constraint limits the optimizer. Clearly, and consistently with figure 7.21,
a vast majority of flight points are sized by the roll maneuver. It can be noticed that the maximal
value of η opt with respect to flight points is around 1.4: more precisely it is equal to 1.437.
Of course η > 1 is not physically admissible: it only indicates that, according to this method,
the aircraft is undersized given the specified constraints. Also appealing on the graph is the clear
visualization of the sizing points: η opt decreases with the Mach number. As expected from
a handling qualities perspective, sizing cases for control surfaces are located for low dynamic
pressure. As already discussed these are also the cases with maximum longitudinal instability, yet
it does not seem to enter into account here. However interestingly we note that a few low-speed
sizing cases are sized no more by the lateral maneuver but by the longitudinal maneuver 4 .

To sum up this part, the optimization principle is promising for it allows viewing rapidly which
flight point is sizing, and what are the associated constraints. However up to this point the flying
wing design seems infeasible from a control perspective. Whether it is an actual design issue, or an
issue linked to the problem formulation will be investigated hereafter, with a method refinement.
3
That this is only first-order is developed in the next section.
4
For sake of clarity in the figure, the distinction between maximum amplitude and maximum rate constraints
was omitted for each constraint.
7.3. Evaluation of Different Methods for Control Surfaces Codesign 179

Control Surfaces Response


Elevons Deflections (in ˚)

δm1
δm2
δm3
δm4
δm5
δn
Time (s)
Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

˙ 1
δm
˙ 2
δm
˙ 3
δm
˙ 4
δm
˙ 5
δm
˙
δn

Time (s)
Aircraft Response

q
q,β,p,r,φ

β
p
nz

r
φ
nz

Time (s)

Figure 7.20: Comparison of temporal responses for η = 1 (dotted lines) and η opt = 0.7415 (solid
lines) computed through decoupled optimization.
180 Chapter 7. Co-Design of Control Surfaces and Laws

90

80

70

60
Sum of limiting constraints

50

40

30

20

10

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
T
˙ i max nzc →δm
δm ˙i ∆n zc

2 opt , K opt ))
∆δmi max T ew →δm i
(P (η

2 opt , K opt ))
˙
δmi max T ˙
ew →δmi (P (η

1 max (P (η opt , K opt ))
T
∆δmi max φc →δmi c φ

1 max (P (η opt , K opt ))
T
˙ i max φc →δmi
δm ˙ φ c

Figure 7.21: Amount of times each constraint is limiting, on the whole flight envelope (156 flight
points), after decoupled optimization. A constraint is considered limiting if its normalized value is
above 0.99. The sum of limiting constraints may exceed the number of flight points, for several
constraints may be active for a single flight point.
7.3. Evaluation of Different Methods for Control Surfaces Codesign 181

Figure 7.22: η opt as a function of flight point, computed with decoupled optimization, together
with associated limiting constraints.

7.3.2 Codesign with LFR Parametrization of the Allocation

Codesign with LFR parametrization of the allocation is explained, then results on the example
flight point are analyzed, and finally this method is applied on the entire flight envelope in order
to find the sizing case for the control surfaces span.

7.3.2.1 Problem Formulation

Previous methode relied on the fact that at first order, control gains may be decoupled from η
optimization. However this lead to infeasible designs. We therefore expect improvements on the
optimum design by performing some “real” codesign, that is simultaneously optimizing control
laws gains and physical parameter η . The procedure is the following one:

Step 1:

γ∞ ← min T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η init ), K) (7.46)


K ∞
such that:
∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η init )
182 Chapter 7. Co-Design of Control Surfaces and Laws

ω02 nref
z + z̃1
s2 +2ξ0 ω0 s+ω02 −
Reference pitch dynamics
1 φref + z2
(1+s/τrp )(1+s/τsp )

Reference roll dynamics
2
ωdr βref + z3
2
s2 +2ξdr ωdr s+ωdr

Reference yaw dynamics

η η α Ve
g zα
nz
q
LFR
δmequi LFR Aircraft β
nzc C ∗ Law Pseudo-inverse 2
ωact u
δmi i = 1..10 Representation
Allocation δn s2 +2ξ
act ωact s+ωact
2 Delay p
φc δlequi Actuators model r
βc Y ∗ Law δnequi ew q 1+ V

3Lz
s wz φ
σz 2L z e
πVe (1+ Lz s)2
Ve

Turbulence model

Figure 7.23: Codesign with parameterized control allocation.

Step 2:

(η opt , K opt ) ← min η (7.47)


η,K

such that:
1
Tn →δmi ∆nzc max (P (η), K) ≤1
∆δmi max zc ∞
1 max
T ˙ ∆nzc (P (η), K) ≤1
δmi max nzc →δmi
˙ ∞
2
Te →δmi (P (η), K) ≤1
∆δmi max w ∞
2
T (P (η), K) ≤1
˙ i max ew →δm˙ i
δm ∞
1
Tφ →δmi φc max (P (η), K) ≤1
∆δmi max c ∞
1 max
T ˙ φc (P (η), K) ≤1
δmi max φc →δmi
˙ ∞
T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η), K) ≤ 1.05 ∗ γ∞

∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η)

A first synthesis aims at obtaining the optimal H∞ value γ∞ . Contrary to previous case,
7.3. Evaluation of Different Methods for Control Surfaces Codesign 183

gains coming from this optimization are not kept. Instead they are computed again in the second
step, to ensure that the optimal norm of the 3x3 transfer T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } remains below

γ∞ , multiplied by a relaxation factor, here 1.05. This relaxation factor ensures that the H∞ norm
of the 3x3 transfer does not get stuck in a possible local optimum of γ∞ versus η .

Please not that the complexity of the second step has increased by an order of magnitude
compared to the decoupled optimization: the constrained optimization problem now features 17
variables (16 control law gains and η ) and 32 constraints. figure 7.23 sums up the procedure.

7.3.2.2 Results on a Single Flight Point

On the flight point M.35, H = 3300f t, m = 300T , following results are obtained:

γ∞ = 0.1042 for η init = 0.2 (7.48)


opt
η = 0.4161 (7.49)

Remark 7.5
For numerical reasons, we found that the codesign optimizer converges better if the initial synthesis
is performed with a small value for η init . This is why we chose to perform the initial synthesis
with η init = 0.2.

One first remarks that the η opt value with codesign optimization has decreased by 43%
compared to the value obtained with decoupled optimization. This is a first hint of the gain in
optimality of coupling the problems versus decoupling. Constraints at the optimum associated
with those optimal values are then summarized on figure 7.25. Once again the constraint on
deflection rates for bank angle input is limiting. However this time the elevon 2 is limiting: the
different η opt value implies a different allocation. Also as η opt is smaller, other constraints
than the limiting one tend to be closer to 1. Even though the optimal point is different, sizing
constraints are consistent from decoupled optimization to codesign.

Then figure 7.24 compares temporal responses for a synthesis with η = 1, and for η opt = 0.4161
with gains computed through codesign. Temporal responses after codesign are similar to those
of the initial synthesis: it shows the ability to recover a desired closed-loop model directly from
a single constraint on a 3x3 transfer. Only the initial response to the bank angle step input is
slightly smaller, which is sufficient to attenuate the initial peak in deflection rate. This is how we
explain such a gain in η opt : by slightly slowing down the initial response to the input of φc ,
deflections are more “spread” in time: for a fixed η this would mean smaller deflections, but as
in our case maximal deflection and rate are fixed, this means a gain in η opt . This is clearly an
optimal point that was not captured by the decoupled optimization, showing the benefit of the
integrated design and control approach.
184 Chapter 7. Co-Design of Control Surfaces and Laws

Control Surfaces Response


Elevons Deflections (in ˚)

δm1
δm2
δm3
δm4
δm5
δn
Time (s)
Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

˙ 1
δm
˙ 2
δm
˙ 3
δm
˙ 4
δm
˙ 5
δm
˙
δn

Time (s)
Aircraft Response

q
q,β,p,r,φ

β
p
nz

r
φ
nz

Time (s)

Figure 7.24: Comparison of temporal responses for η = 1 (dotted lines) and η opt = 0.4161 (solid
lines) computed through codesign with LFR parametrization of the allocation.
7.3. Evaluation of Different Methods for Control Surfaces Codesign 185

0.9

0.8

0.7

0.6
H∞ norms of constraints

0.5

0.4

0.3

0.2

0.1

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
T
˙ i max nzc →δm
δm ˙i ∆n zc

2 opt , K opt ))
∆δmi max T ew →δm i
(P (η

2 opt , K opt ))
˙
δmi max T ˙
ew →δmi (P (η

1 max (P (η opt , K opt ))
T
∆δmi max φc →δmi c φ

1 max (P (η opt , K opt ))
T
˙ i max φc →δmi
δm ˙ φ c

Figure 7.25: Bar of constraints values after codesign with parameterized allocation, η opt = 0.4161.

7.3.2.3 Codesign with Parameterized Allocation on the Entire Flight Envelope

We have shown the benefits of the integrated plant/controller optimization approach on a single
flight point, now let us see what happens on the entire flight envelope. figure 7.26 shows the
surface η opt = f (M , H) and sizing constraints associated with each flight point. A first appealing
fact is the maximum value of η opt over the whole flight envelope: it is now located a bit above
186 Chapter 7. Co-Design of Control Surfaces and Laws

0.6. Precisely the maximum is 0.6269. It means changing the optimization method also changed
the conclusion concerning the BWB design from a control perspective: the initial design η = 1
now becomes feasible, it even becomes over-sized. Yet this figure also offers a lot of similarities
with the previously plotted curve with decoupled optimization: clearly η opt is smoothly decreasing
as Mach number increases. Even though it is not as clear on the figure, an increase of η opt when
the altitude increases for a given Mach number can be observed. This is consistent with the loss
of dynamic pressure which makes the controls less effective, and this gives us confidence in the
fact that this procedure indeed captures the physical phenomena at stake for sizing.

Figure 7.26: η opt as a function of flight point, computed with codesign with parameterized
allocation.

Once again most cases are sized by the lateral maneuver constraint. However interestingly in
some high-altitude corners of the flight envelope the turbulence constraint becomes sizing. This
observation is verified by the analysis of figure 7.27, where the sum of limiting constraints (that is
superior to 0.99) for all flight points is displayed for each control surface. Once again the maximum
deflection rate in response to bank order represents a vast majority of sizing cases, particularly for
controls 2 and 3. Summing up all these bars gives 227: 45% of flight points are sized by more
than one criterion. This is a significant improvement over the decoupled optimization results.

As a conclusion for this method, remarkable improvements over the decoupled optimization
were achieved by optimizing together the compensator and physical parameter η . The initial
design is now proved feasible, and even over-sized. Roll maneuver is shown to be sizing in most
cases. However the allocation choice constrains only one control surface to be sizing for each
flight point. This issue is solved in next section.
7.3. Evaluation of Different Methods for Control Surfaces Codesign 187

120

100

80
Sum of limiting constraints

60

40

20

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
T
˙ i max nzc →δm
δm ˙i ∆n zc

2 opt , K opt ))
∆δmi max T ew →δm i
(P (η

2 opt , K opt ))
˙
δmi max T ˙
ew →δmi (P (η

1 max (P (η opt , K opt ))
T
∆δmi max φc →δmi c φ

1 max (P (η opt , K opt ))
T
˙ i max φc →δmi
δm ˙ φ c

Figure 7.27: Amount of times each constraint is limiting, on the whole flight envelope (156 flight
points), with parameterized codesign. A constraint is considered limiting if its normalized value is
above 0.99. The sum of limiting constraints may exceed the number of flight points, for several
constraints may be active for a single flight point.
188 Chapter 7. Co-Design of Control Surfaces and Laws

ω02 nref
z + z̃1
s2 +2ξ0 ω0 s+ω02 −
Reference pitch dynamics
1 φref + z2
(1+s/τrp )(1+s/τsp )

Reference roll dynamics
2
ωdr βref + z3
2
s2 +2ξdr ωdr s+ωdr

Reference yaw dynamics

η α Ve
g zα
nz
q
δmequi LFR Aircraft β
C ∗ Law 2
ωact
nzc Kalloc δmi i = 1..10
2
u Representation
Delay p
δn s2 +2ξ act ωact s+ωact
φc δlequi Actuators model r
βc Y ∗ Law δnequi ew q 1+ V

3Lz
s wz φ
σz 2L z e
πVe (1+ Lz s)2
Ve

Turbulence model

Figure 7.28: Codesign with tunable control allocation.

7.3.3 Codesign with Tunable Allocation

A last alternative of the codesign procedure is presented now, where the allocation block becomes
now tunable. The principle of the method is explained, then an example optimization illustrates
the approach, and finally results on the entire flight envelope are analyzed.

7.3.3.1 Problem Formulation

While the codesign approach revealed powerful, allowing a handling qualities sizing directly in
closed-loop, sizing cases for a given flight point are limited by the allocation priority order. As
a consequence with previously described method the majority of flight points were sized by only
one criterion. We believe there is room for improvement on this side: instead of using a fixed
allocation, parameterized by η , the allocation is now a tunable parameter of the optimization.
Tunable blocks for this synthesis are represented on figure 7.28. Mathematically the problem
writes:

Step 1:

γ∞ ← min T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η init ), K, Kalloc


init
) (7.50)
K ∞
such that:
∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η init )
7.3. Evaluation of Different Methods for Control Surfaces Codesign 189

Step 2:

(η opt , K opt , Kalloc ) ← min η (7.51)


η,K,Kalloc

such that:
1
Tn →δmi ∆nzc max (P (η), K, Kalloc ) ≤1
∆δmi max zc ∞
1
T ∆nzc max (P (η), K, Kalloc ) ≤1
˙ i max nzc →δm˙ i
δm ∞
2
Te →δmi (P (η), K, Kalloc ) ≤1
∆δmi max w ∞
2
T ˙ (P (η), K, Kalloc ) ≤1
δmi max ew →δmi
˙ ∞
1
Tφ →δmi φc max (P (η), K, Kalloc ) ≤1
∆δmi max c ∞
1
T φ max (P (η), K, Kalloc ) ≤1
˙ i max φc →δm˙ i c
δm ∞
T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η), K, Kalloc ) ≤ 1.05 ∗ γ∞

∀p ∈ C, p pole of P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
K internally stabilizes P (η)

The process is very similar to that presented in section 7.3.2, except Kalloc is now a variable
for the optimizer. Initially Kalloc is a 11x3 matrix. However we have knowledge on the expected
shape of this matrix: we want symmetrical deflections for pitch, anti-symmetrical deflections for
roll and yaw with the rudder only. Therefore we constrain the structure of Kalloc , in order to
limit the space of solutions for the optimizer:
pitch roll

Kalloc Kalloc 0
 pitch roll
Kalloc = Kalloc −Kalloc 0  (7.52)

yaw
0 0 Kalloc

Thanks to this manipulation Kalloc now has only 11 degrees of freedom instead of 33
initially. As a consequence the second step is a constrained optimization with a single objective,
32 constraints and 28 (16+1+11) variables.
190 Chapter 7. Co-Design of Control Surfaces and Laws

Control Surfaces Response


Elevons Deflections (in ˚)

δm1
δm2
δm3
δm4
δm5
δn
Time (s)
Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

˙ 1
δm
˙ 2
δm
˙ 3
δm
˙ 4
δm
˙ 5
δm
˙
δn

Time (s)
Aircraft Response

q
q,β,p,r,φ

β
p
nz

r
φ
nz

Time (s)

Figure 7.29: Comparison of temporal responses for η = 1 (dotted lines) and η opt = 0.3885 (solid
lines) computed through codesign with tunable control allocation.
7.3. Evaluation of Different Methods for Control Surfaces Codesign 191

7.3.3.2 Results on a Single Flight Point

Codesign with tunable allocation applied to the flight point M.35, H = 3300f t, m = 300T gives
following results:

γ∞ = 0.1042 for η init = 0.2 (7.53)


opt
η = 0.3885 (7.54)

A 16% decrease is achieved on η opt compared to the codesign with parameterized allocation.
A look at figure 7.30, gathering handling qualities constraints at the optimum, indicates where
this gain comes from. Instead of having only one control surface saturated, as it was previously
the case, all five controls are now saturated for the criterion of deflection rate in response to a
bank angle order. This is confirmed by looking at Kalloc after optimization:

 
0.023 0.3811 0
0.0116 0.3811 0 
 
 
 0.000 0.3811 0 
 
 0.000 0.3811 0 
 
 0.000 0.3811 0 
 
opt  
Kalloc =
 0.023 −0.3811 0  (7.55)
0.0116

−0.3811 0 
−0.3811
 
 0.000 0 
 
 0.000
 −0.3811 0 
 0.000 −0.3811 0 
 

0.000 0.000 0.006

As the roll constraint is limiting, the optimal allocation is to distribute equally among all
controls the roll order. As a consequence all roll criteria are equally saturated. Pitch being
not limiting on this flight point, only two controls are used for pitch control. Such strategy
can also be seen on temporal response of figure 7.29: all five controls are equally used for roll
maneuver, resulting in equal deflection rates for all control. Temporal response of output is similar
to the parameterized codesign: the roll rate is a bit slower to initiate, which allows to gain some
optimality.

7.3.3.3 Codesign with Tunable Allocation on the Entire Flight Envelope

The codesign with tunable allocation method is now applied to the whole flight domain. figure 7.31
shows resulting surface η opt = f (M , H). An improvement of the maximum value of η opt =
0.5996 is found, which is a 4.3% improvement over the sizing value obtained by codesign with
parameterized allocation. Interestingly, sizing flight points, which are once again located at low
Mach number and high altitude, are now sized by a combination of turbulence, longitudinal and
lateral maneuver. As initially expected, strong stabilization to an exogenous input is definitely
192 Chapter 7. Co-Design of Control Surfaces and Laws

0.9

0.8

0.7

0.6
H∞ norms of constraints

0.5

0.4

0.3

0.2

0.1

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
T
˙ i max nzc →δm
δm ˙i ∆n zc

2 opt , K opt ))
∆δmi max T ew →δm i
(P (η

2 opt , K opt ))
˙
δmi max T ˙
ew →δmi (P (η

1 max (P (η opt , K opt ))
T
∆δmi max φc →δmi c φ

1 max (P (η opt , K opt ))
T
˙ i max φc →δmi
δm ˙ φ c

Figure 7.30: Bar of constraints values after codesign with tunable allocation, η opt = 0.3885.

sizing for some specific flight points. Moreover these results evidence the importance of the
control allocation module for a correct sizing of the aircraft: it has a direct influence on the
design.

Finally the sum of limiting constraints for each flight point is displayed on figure 7.32. The
roll constraint on deflection rates remains sizing in the vast majority of cases, accordingly with
7.3. Evaluation of Different Methods for Control Surfaces Codesign 193

Figure 7.31: η opt as a function of flight point, computed with codesign with tunable allocation.

figure 7.31. However the roll deflection constraint becomes also sizing for all control surfaces for
nearly 60 flight points. More interestingly, the turbulence constraint sizes the control surfaces in
nearly 20 flight points. In average, each flight point is now sized by 7 criteria in average: this
explains the remarkable improvements obtained on η opt whatever the flight points.

As a conclusion for this section, results on all flight points of the three optimization methods
are displayed together on figure 7.33. Going from decoupled to coupled optimization clearly shows
a significant improvement in all cases. Then adding the allocation as a variable parameter still
improves the design every time.

Conclusion

Three different methods were investigated in order to size the BWB control surfaces directly
with closed-loop handling qualities criteria. We have shown that while parameterizing the control
allocation module with the sizing physical parameter may at first sight decouple the control
problem from the sizing problem, this approach misses the optimal sizing points and then leads
to erroneous conclusions concerning the sizing of the aircraft. On the contrary, combining the
control and sizing problems into a single optimization allows finding more optimal solutions. We
have also shown the importance of the control allocation method for the design: restricting to
pseudo-inverse solutions may limit the capabilities of the aircraft, whereas a “flattened” allocation
allows downsizing the aircraft from a control perspective. Also from a sizing perspective, we now
have evidence that the studied configuration is adequately controlled on all three axes, on all the
194 Chapter 7. Co-Design of Control Surfaces and Laws

140

120

100
Sum of limiting constraints

80

60

40

20

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
T
˙ i max nzc →δm
δm ˙i ∆n zc

2 opt , K opt ))
∆δmi max T ew →δm i
(P (η

2 opt , K opt ))
˙
δmi max T ˙
ew →δmi (P (η

1 max (P (η opt , K opt ))
T
∆δmi max φc →δmi c φ

1 max (P (η opt , K opt ))
T
˙ i max φc →δmi
δm ˙ φ c

Figure 7.32: Amount of times each constraint is limiting, on the whole flight envelope (156 flight
points), with tunable allocation codesign. A constraint is considered limiting if its normalized
value is above 0.99. The sum of limiting constraints may exceed the number of flight points, for
several constraints may be active for a single flight point.

flight envelope, with the initial control surfaces configuration, with normal law and full control
authority, with reasonable hypotheses on the actuators, sensors and delays chain. Consistently
with the process proposed in chapter 4 the next step in order to demonstrate the viability of this
7.3. Evaluation of Different Methods for Control Surfaces Codesign 195

decoupled optimization
codesign with parameterized allocation
codesign with tunable allocation
η opt

Flight Point number

Figure 7.33: Comparison of η opt with three optimization methods, for all flight points.

aircraft configuration would be to examine the remaining control authority facing different failure
scenarios.

From a control perspective, the nonsmooth H∞ synthesis we used proved very efficient in
finding satisfactory solutions to the control problem of three-axes synthesis, and more impor-
tantly to the codesign problems. However solving sizing problems is not the initial purpose of
these algorithms: as a consequence the user has to find tricks in order to specify sizing objec-
tives and constraints under the form of norms of transfer functions. A perspective would be to
extend those tools in order to handle more diverse problems, such as variables minimization, or
temporal criteria. If going on this direction, simulation-based design could be explored, together
with optimization algorithms more general and dedicated to non-linear problems, such as genetic
algorithms. However this would probably be at the expense of an increase in computational time.
CONCLUSION AND PERSPECTIVES

197
CONCLUSION AND PERSPECTIVES 199

Summary of Contributions

The main contribution of this thesis is to formulate and solve the problem of sizing control surfaces
and controlling a blended wing-body aircraft in closed-loop, using mathematical and numerical
tools coming from the control community. How this was achieved is summarized hereafter.

After describing the state of the art and various challenges of the BWB configuration in
chapters 1 and 2, chapter 3 was dedicated to formalizing the problem of integrated design and
control on a simplified example, namely the inverted pendulum. Even though this example may
look academical, an original contribution of ours consisted in formalizing the interactions between
physical parameters, namely stick size and bandwidth, and control laws gains. Most importantly
we formulated the integrated design and control problem under a novel way of a constrained
optimization problem, with sizing criteria formulated as constraints on frequency norms. This
problem was then solved using nonsmooth optimization algorithms for controllers tuning. Basically
this chapter contains the essence of our work on its simplest form: and being able to capture the
principle of the problem, and to deliver it in a pedagogical form to the reader, is in our opinion a
significant achievement.

Then chapter 4 formalizes in a general way the different stages of sizing an aircraft which
is possibly unstable, from a perspective of handling qualities at future projects phase. Indeed,
segregation among different disciplines such as aerodynamics, open- and closed-loop handling
qualities, flight control systems, which physical parameters they respectively influence, interactions
between these disciplines, and spreading of these disciplines on a time scale, had not been properly
formalized to our knowledge. This is especially true for the specific problem of an unconventional
and possibly unstable aircraft. Having this process in mind is from our point of view fundamental
in order to segregate the multiple challenges the designer faces when sizing an aircraft departing
from the usual design and knowledge. A second significant contribution of this chapter is then
to integrate the codesign approach we developed into this process of unstable aircraft sizing: we
show that the codesign is an improvement over the traditional process, which shortens iterations
and leads to a more optimal design. Once again, we believe it clarifies the field of application
of our method in aircraft preliminary design. In fact two alternatives to the iterative process are
proposed: the first one features the integrated design of actuators bandwidth and control, later
worked out in chapter 5, and the second one is about integrated design of control surfaces size
and control, developed extensively in chapter 7.

The methodology proposed in chapter 4 and developed on an example in chapter 3 is then


applied to the problem of aircraft longitudinal stabilization. The main achievement of this chapter
is to formulate the problem of actuators bandwidth sizing under the form of a H2 /H∞ opti-
mization problem, with constraint on the closed-loop transfer to ensure adequate performance.
This formulation allows finding the minimal actuator bandwidth that guarantees a certain level of
performance. A second novelty of this approach is to include control surfaces hinge moments in
the optimization criterion. As a consequence the optimization criterion, even though not being ex-
actly the actuators power consumption, is strictly monotonous with respect to power consumption.
Using this approach we found that the optimal solution from the actuators power consumption
perspective is to stabilize the aircraft using only the smallest control surface; moreover reasonable
200 CONCLUSION AND PERSPECTIVES

bandwidth, namely below 2.5Hz, is sufficient for controlling the instability. Such a result, and the
way it was obtained, is original to our knowledge. However a major limitation of the approach is
the absence of guarantee of any maximal control level for stabilization. Also elevons should be
sized not only with respect to longitudinal criteria, but also lateral.

To overcome those limitations while retaining the principle of the approach, an alternative
study is proposed. It aims at conjointly optimizing outer control surfaces total span together
with longitudinal and lateral flight control laws, taking into account industrial requirements in
the control law structure. For that purpose, a first step requires the computation of models
continuously parameterized by the parameter to size. Chapter 6 explains how discretized state-
space representations are first obtained from light CFD computations, corrected with more
accurate aerodynamic data, and then how these models are approximated into a continuous LFR
representation using a least-square approximation. Parameterizing the state-space representations
not by uncertain or varying flight parameters, but by the physical size of control surfaces in order
to get a model suitable for further optimization using optimizers for controllers tuning, is the main
original contribution of this part.

Chapter 7 represents the main original work of our study: it is based on the rationale and
conclusions of all preceding parts. Parameterized state-space representations obtained in chapter 6
are optimized under constraints on closed-loop performance and handling qualities constraints. A
frequency-domain approach is used in order to set these handling qualities constraints, written as
maximum amplitude and rate of controls in response to exogenous inputs or disturbance: while
this approach has already been developed in literature, we propose a discussion on the meaning
and limitations of these criteria in the temporal domain, which should be helpful to the engineer
willing to apply our techniques for aircraft sizing. Writing the problem of C ∗ /Y ∗ gains synthesis
under the form of a single constraint satisfaction problem is also a significant step and an enabler
for optimizations routines taking into account prescribed structures for controllers. Finally once
these constraints are properly written, three alternative optimization procedures are proposed in
order to solve the problem of gains and physical parameters computation, each one differing from
the others by the variables provided to the optimizer. A first synthesis considers a decoupled and
sequential optimization of the flight control law first, and the control surfaces span afterwards,
taking advantage of the control allocation module being parameterized by the control surfaces span
parameter. While we show that this approach leads to suboptimal design, we still find an interest in
presenting it from an industrial application perspective: indeed this approach does not specifically
require a gains optimization through H∞ synthesis: engineers may apply any synthesis techniques
they are used to, and optimize the physical parameters with fixed gains in a second step. Yet
significant improvements in the design are achieved when applying the codesign technique, that
is a coupled optimization of flight control laws and physical parameters, instead of the sequential
approach. The initial control surfaces layout now proves viable, with even margin on control
surfaces size with respect to the specified handling qualities criteria. We show that the outer
control surfaces span may be reduced by 45% before the constraints are saturated. Interestingly,
most flight points in the flight envelope turn out to be sized by a roll criterion; however the
most sizing points, at low Mach number and high altitude, are also sized by turbulence rejection,
which implicitly means instability control sizes the aircraft. Finally a serious dependency of the
optimized design versus the control allocation method is shown: including the allocation into the
CONCLUSION AND PERSPECTIVES 201

optimization then leads to a few percents design improvements.

As a conclusion for our work, we have achieved in finding a technique to size an aircraft
directly from closed-loop specifications, both in longitudinal and lateral: this technique is capable
of delivering relevant information on the sizing point over the flight envelope, and sizing handling
qualities criteria. This should be helpful for the designers to make decisions on the next generation
of aircraft.

Perspectives

This work opens quite a diverse outlook. First the codesign approach, extensively developed
here for the sizing of a BWB , would in our opinion provide striking answers to the design of
other unconventional configurations. For instance we could imagine reducing the VTP size of
an aircraft up to the point where it becomes yaw-unstable, at the expense of controlling this
unstable mode with differential thrust of the engines. This would straightforwardly translate into
the following codesign problem: what is the minimum VTP size which guarantees a certain level
of performance in closed-loop, while preserving a maximum amplitude on rudder deflections and
thrust increase when facing a calibrated exogenous disturbance such as lateral turbulence?

However even though the BWB is the perfect aircraft for applying our proposed techniques, it
has a drawback: we have no vision on a correctly designed civil BWB, for none has ever flown. As
a consequence, our conclusions cannot be compared to more traditional approaches, so it is hard
to build confidence on our process and the proposed formulation for handling qualities constraints.
Therefore a quick way to gain confidence in our approach would be to apply it on a conventional
aircraft, and to compare it to classical handling qualities sizing. The difference in sizing results
would probably imply an improvement of the handling qualities criteria compared to the ones we
proposed.

More generally a lot of work would deserve to be done on the handling qualities criteria
definition. As already emphasized, writing these as frequency norms certainly has advantages from
an optimization point of view, yet it definitely lacks physical interpretation. A next ambitious step
could be to include temporal criteria. Yet this would imply changing the optimization methods, and
the problem formulation. Using more general optimization algorithms, such as genetic algorithms,
could reveal powerful on our problem. It is often argued in the control community that controllers
synthesis with this type of general nonlinear algorithms is very demanding from a computational
time perspective, which is probably true, however it would make sense for our integrated design
and control problem. It would allow more diverse criteria, such as simulation-based design, and
a more flexible choice in the definition of the objectives. Nonsmooth optimization routines for
controllers tuning, while very efficient in finding solutions for complex problems such as codesign,
suffer today from limitations due to the fact that they are initially not intended for that purpose.
More precisely, as a designer and not only a control designer, one would want to specify criteria and
objectives not only under the form of norms of transfer functions5 . Going towards more versatile
5
It turns out recent versions of systune allow some kind of temporal specifications, yet we have not tried it. From
202 CONCLUSION AND PERSPECTIVES

optimization algorithms would also allow for doing some “real” multiobjective optimization, as
well as giving relevant information about sizing points: as a designer, a Pareto front of several
objectives is certainly more relevant than a single sizing point, whatever optimal it may be. Such
information is not straightforward to obtain in the tools we used.

To go on concerning multiobjective optimization, a major difficulty encountered during this


work appeared comparing things that simply cannot be compared: for instance physical measures
such as power consumption or elevon deflection, with H2 or H∞ norms. Control designers usu-
ally do this everyday without trouble; yet for instance H2 and H∞ norms have different units
in general. When physical design is considered, that is when performing codesign, comparing
values with different units definitely becomes shocking. Normalizing is the common answer to
that concern, and this is what we did when writing the different constraints, however it is not
completely satisfactory. Indeed the choice of the figures for normalization already induces some
design choice: should we normalize by the initial value? By the expected optimum? A “poten-
tiometer” systematically arises, which is exactly what we tried to avoid throughout this report. On
the same order of ideas, a common solution consists in adding weights to the different channels or
objectives. This manual tuning is very widespread in literature ( see for instance [Sahasrabudhe
et al., 1997]), yet once again it turns into a hand-tuning by the designer. These questions have
however been studied extensively by the multiobjective optimization community; methods coming
from this community, such as game theory, could be worth investigating.

Multidisciplinary Analysis and Optimization (MDAO) is also another attractive research field.
Taking advantage of recent progress in numerical tools for optimization, this field aims at integrat-
ing together different disciplines, typically Aerodynamics, Structure and Engines design for aircraft
design (see for instance [Prigent et al., 2013]). Our problem would translate in this framework
into the inclusion of a flight control discipline in an aircraft sizing and optimization process.

From a theoretical standpoint the use of efficient optimization routines for controller tuning
has opened again an old topic in control theory: are requirements for a control law feasible, and if
not, what is the maximum achievable level of performance? Indeed optimization algorithms have
proved very efficient in finding the best solutions for a feasible set of constraints. The control
designer task has now evolved from control law tuning to specifications tuning. Certificates of
emptiness of solutions for given constraints is a hard problem, yet decision tools indicating to the
designer if the constraints are realizable is an appealing field of research. Similarly, the problem of
finding the minimal control law order to achieve a certain level of performance is to our knowledge
still open.

To come back to BWB design, we have shown with a good level of confidence that controlling
the aircraft is feasible in normal law, that is provided all sensors, measurements and actuators are
available. Consistently with the process developed in chapter 4, the next step would now consists
in a comprehensive flight control systems study. It means making assumptions on the actuators
technology, energy distribution, systems architecture, and then analyzing the capabilities of the
aircraft under different outage scenarios. Such failure cases, intentionally kept for future work

our information it also appears that these temporal specifications are in fact turned into frequency constraints. What
these routines still lack for becoming essential for codesign solving is the ability to simply specify any objective.
CONCLUSION AND PERSPECTIVES 203

for they deserve a dedicated study, will in the end certainly reveal more sizing than the handling
qualities constraints examined here. The coupled architecture of the BWB will once again be
a challenge and an opportunity for the combinatorial exploration of all failure cases: a challenge
because any outage (for instance actuator jamming) will imply degraded performance on all axes,
making the analysis much more difficult compared to actual practices, and an opportunity because
the redundancy of control surfaces will make the architecture much more resilient to failure.
APPENDIX

205
Appendix A

Analytical Expression of State-Space


Representations

Contents
A.1 Longitudinal Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . 207

A.2 Lateral Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208

A.1 Longitudinal Coefficients

Developing coefficients in state-space matrices of section 2.1.4.5 gives:

−ρVe SCx ∂F
xV = + , xα = −2gki Czα ,
m ∂Ve
−2glki
xq = Czq , xθ = −g,
Ve
−2g ρVe S
zV = , zα = Czα ,
Ve 2 2m
ρSl ρVe 2 Sl
zq = Czq , mα = CmGα,
2m 2mRyy
ρVe Sl2 1 ∂F
mq = CmG q , xδx = ,
2mRyy m ∂δx
ρVe S
xδmi = −2gki Czδmi , zδmi = Czδmi ,
2m
ρVe 2 Sl
mδmi = CmG
δmi
2mRyy

207
208 Appendix A. Analytical Expression of State-Space Representations

A.2 Lateral Coefficients

Developing coefficients in state-space matrices of section 2.1.5.5 gives:

ρVe S ρSL
yβ = Cyβ , yp = Cyβ
2m 2m
ρSL ρVe S
yr = Cyβ , yδmi = Cyδmi
2m 2m
ρVe S ρVe 2 SL
yδn = Cyδn , łβ = Cl
2m 2mRxx β
ρVe 2 SL ρVe 2 SL
lp = Cl , lr = Clr
2mRxx p 2mRxx
ρVe 2 SL ρVe 2 SL
lδn = Clδn , nβ = Cnβ
2mRxx 2mRzz
ρVe 2 SL ρVe 2 SL
np = Cnp , nr = Cnr
2mRzz 2mRzz
ρVe 2 SL ρVe 2 SL
nδmi = Cnδmi , nδn = Cnδn
2mRzz 2mRzz
RESUME ETENDU EN FRANCAIS

209
Annexe B

Dimensionnement Conjoint de
Surfaces de Contrôle et Lois de
Commande pour Configurations
d’Avions Non-Conventionnelles

Cette annexe a pour but de fournir un résumé étendu en français du travail de thèse décrit
précédemment dans la langue de Shakespeare. Néanmoins cette partie, bien que se voulant
auto-complète, ne saurait prétendre à l’exhaustivité. Le lecteur désirant avoir plus de détails
sur le contenu technique du travail réalisé durant cette thèse est donc invité à se diriger vers les
parties précédentes. Dans ce résumé étendu, deux points principaux correspondant aux deux
contributions principales de la thèse seront abordés. Le premier concerne la formulation d’un
processus de dimensionnement d’un avion instable. C’est une redite partielle du chapitre 4. Le
second concerne la synthèse de lois trois-axes avec le formalisme H∞ , et surtout l’application
de ce formalisme au problème du dimensionnement conjoint de tailles de gouvernes et lois de
commande, le fameux codesign familier au lecteur.

Contents
B.1 Description d’un processus de dimensionnement d’un avion instable
au stade avant-projets . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
B.1.1 Processus itératif . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
B.1.2 Processus couplé pour le dimensionnement des gouvernes . . . . . . 214
B.2 Obtention d’un modèle aérodynamique paramétré . . . . . . . . . . . . 216
B.2.1 Paramétrisation géométrique des gouvernes de bord de fuite . . . . . 216
B.2.2 Calcul de modèles aérodynamiques pour des envergures de gouvernes
discrétisées . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
B.2.3 Approximation polynômiale et représentation LFR des efficacités aé-
rodynamiques de gouvernes . . . . . . . . . . . . . . . . . . . . . . 219
B.3 Codesign de gouvernes et lois de commandes sous contraintes qualités
de vol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
B.3.1 Structure des lois de commande . . . . . . . . . . . . . . . . . . . . 221
B.3.2 Modèle d’allocation . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
B.3.3 Synthèse simultanée de lois 3 axes . . . . . . . . . . . . . . . . . . . 222

211
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
212 Commande
B.3.4 Définition du problème d’optimisation . . . . . . . . . . . . . . . . . 226
B.4 Résultats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
B.4.1 Codesign sur un seul point de vol . . . . . . . . . . . . . . . . . . . 229
B.4.2 Résultats sur l’ensemble du domaine de vol . . . . . . . . . . . . . . 229

B.1 Description d’un processus de dimensionnement d’un avion


instable au stade avant-projets

Le dimensionnement qualités de vol au stade avant-projets n’est pas adapté aux configurations
naturellement instable. Les méthodes de dimensionnement traditionnelles, telles que décrites dans
Roskam [1995], sont notamment basées sur les notions de coefficient de volume, ainsi que sur des
calculs d’équilibre sur l’ensemble du domaine de vol, et des manœuvres temporelles dynamiques,
parfois approximées par des pseudo-équilibres. Pourtant les avions récents sont conçus avec des
marges de plus en plus réduites vis-à-vis de l’instabilité, comparé aux précédentes générations.
L’hypothèse sous-jacente est que lors de la phase de conception détaillée, les ingénieurs chargés de
la conception des lois de commande seront capables de contrôler l’avion avec un comportement
adéquat et des amplitudes de commande respectables, et ce quelle que soit la configuration
géométrique de l’avion. Cette hypothèse est justifiable tant que l’on se limite à concevoir des avions
de forme conventionnelle, historiquement adaptés à un pilotage manuel. En revanche d’un point de
vue méthodologique ce processus de dimensionnement n’est pas satisfaisant : si l’objectif en boucle
ouverte est trop ambitieux du point de vue de la stabilisation, alors un coûteux redimensionnement
devient nécessaire. A l’inverse, si l’objectif se révèle trop conservatif, alors un meilleur optimum
d’un point de vue global aurait pu être atteint. La seule méthode satisfaisante consiste alors à
inclure des lois de commande directement dans le dimensionnement de l’avion.

La section B.1.1 de ce chapitre explique en détail un processus itératif pour pré-dimensionner


d’un point de vue qualités de vol un avion tout en incluant des lois de commande. Par pré-
dimensionnement, il est entendu que c’est à la phase de dimensionnement préliminaire, ou avant-
projets, que nous nous intéressons. En sortie de ce processus, l’avion n’est à l’évidence pas encore
bon pour vol. En revanche on sait alors que la configuration est saine et dé-risquée, et ouvre la
voie à l’étape de conception détaillée. Les limites de ce processus sont également mises en évi-
dence. A partir de ces limites, on définira quelques alternatives à ce schéma de dimensionnement ;
l’alternative la plus prometteuse, appelée “co-design”, sera développée dans la section B.1.2.

B.1.1 Processus itératif

La procédure traditionnelle, dite “itérative”, est présentée en figure B.1. Pour des raisons de clarté,
toutes les contraintes et variables ne sont pas représentées. On ne mentionnera que les contraintes
les plus pertinentes vis-à-vis de notre problème, afin de faciliter la lecture et la compréhension. La
procédure est représentée par échelle de temps croissante du haut vers le bas. Certaines contraintes
sont exprimées spécifiquement à propos du du dimensionnement de l’aile volante BWB , mais
sont adaptables au cas plus général de dimensionnement d’un avion instable.
B.1. Description d’un processus de dimensionnement d’un avion instable au stade
avant-projets 213

Modèle Aérodynamique
Performance (CoP) Définition de la Gouvernes
forme planeur Aire &
Instabilité (Forme en plan + profils)
(Marge Statique)
Empennages

Qualités de Vol
Qualités de Vol Dynamique Actionneurs
Echelle du Temps

Déflection Max Boucle Ouverte : (Bande Passante)


(ou plage de CG) Qualités de Vol
Equilibres Objectifs Lois de Commande
Boucle Fermée : (Pulsation, Amortissement
Lois de Commande Découplage...)
Perturbation
Plage de CG / Manoeuvre
( ou Déflection Max) Vitesse de
Déflection Max
Débattement Max
Pré-Dimensionnement Calcul des
Actionneurs Déflection Max Moments de
Charnière

Moments de Charnière
Max

Estimation Estimation de
Techno Actionneurs : Techno Actionneurs :
EHA, EMA... de Masse Consommation EHA, EMA...
Actionneurs de Puissance
Actionneurs

Dimensionnement Systèmes
Techno Actionneurs :
EHA, EMA... Cas de
Techno Actionneurs : Panne
Risque / Probabilité

Architecture Systèmes :
Découpage des
Gouvernes

Figure B.1 : Processus itératif. L’échelle temporelle est représentée croissante de haut en bas.
Chaque boîte représente une discipline du processus. Les entrées de chaque disciplines sont re-
présentées par des flèches arrivant par le haut de chaque boîte ; les résultats de chaque discipline
sont les flèches provenant du bas de chaque boîte. Quant aux contraintes spécifiques à chaque
discipline, proviennent du côté.
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
214 Commande
B.1.2 Processus couplé pour le dimensionnement des gouvernes

L’inconvénient majeur de la procédure de dimensionnement présentée précédemment est son coût


en termes de temps de développement si un redimensionnement des gouvernes, des empennages
ou de la forme planeur s’avère nécessaire suite à l’analyse dite “qualités de vol en boucle fermée”.
Plus précisément, si l’évaluation du critère qualités de vol en boucle fermée indique des amplitudes
de commande inacceptables, le modèle géométrique doit alors être modifié. Le modèle aérody-
namique ainsi que les lois de commande doivent alors également être mis à jour. En particulier
ce dernier point n’est pas nécessairement trivial : si une partie des dernières avancées en théorie
de l’automatique vise à automatiser autant que possible la synthèse de lois, une procédure tota-
lement automatique n’est pas forcément disponible quel que soit le modèle aérodynamique. De
plus nous n’avons aucune garantie que le modèle géométrique mis à jour impliquera des niveaux
de commande acceptables.

Pour ces raisons nous proposons une amélioration de la procédure itérative actuelle : la procé-
dure dite de co-design pour dimensionnement de gouvernes, présentée en figure B.2. La différence
principale entre cette procédure et celle présentée précédemment provient de la définition du critère
de qualités de vol boucle fermée.

Désormais les débattements et vitesses de débattements maximaux ne sont plus des


sorties du critères, mais deviennent des contraintes.

Ceci signifie que désormais nous imposons un niveau maximal de déflections, typiquement
imposé par des contraintes géométriques de butée (aux alentours de ±30˚ sur les avions conven-
tionnels) ainsi que de vitesse de débattement, provenant de contraintes technologiques sur les
actionneurs.

Il est ensuite nécessaire de construire un modèle géométrique et aérodynamique paramétré


par les grandeurs que l’on souhaite dimensionner. Dans notre application on recherche la taille
minimale de gouvernes qui satisfait les contraintes de qualités de vol en boucle fermée, tout en
respectant les contraintes d’amplitude maximale de commande. La construction de ce modèle
paramétré permet un couplage entre le dimensionnement géométrique, la synthèse de lois de
commande et l’évaluation des qualités de vol en boucle fermée. Par conséquent en une seule étape
on garantit un dimensionnement optimal des gouvernes tout en satisfaisant des contraintes de
qualités de vol en boucle fermée.

Cette procédure alternative est philosophiquement identique au processus de code-


sign de dimensionnement conjoint de la loi de commande stabilisante et de la longueur
du pendule décrit dans la section 3.5.2. C’est donc une justification a posteriori du
développement de cet exemple académique. L’application de cette procédure au dimen-
sionnement des élevons sur l’aile volante est décrite ci-dessous.
B.1. Description d’un processus de dimensionnement d’un avion instable au stade
avant-projets 215

Modèle Aérodynamique
Performance (CoP) Définition de la Gouvernes
forme planeur Aire &
Instabilité (Forme en plan + profils)
(Marge Statique)
Empennages

Taille Minimale des Surfaces

Qualités de Vol
Qualités de Vol Dynamique Actionneurs
Echelle du Temps

Déflection Max Boucle Ouverte : (Bande Passante)


(ou plage de CG) Qualités de Vol
Equilibres Objectifs Lois de Commande
Boucle Fermée : (Pulsation, Amortissement
Lois de Commande Découplage...)
Perturbation
Plage de CG / Manoeuvre
( ou Déflection Max) Déflection Max et
Vitesse de
Débattement Max
Pré-Dimensionnement Calcul des
Actionneurs Déflection Max Moments de
Charnière

Moments de Charnière
Max

Estimation Estimation de
Techno Actionneurs : Techno Actionneurs :
EHA, EMA... de Masse Consommation EHA, EMA...
Actionneurs de Puissance
Actionneurs

Dimensionnement Systèmes
Techno Actionneurs :
EHA, EMA... Cas de
Techno Actionneurs : Panne
Risque / Probabilité

Architecture Systèmes :
Découpage des
Gouvernes

Figure B.2 : Processus couplé pour le dimensionnement des gouvernes. La différence principale
entre cette figure et celle du processus dit “itératif” réside dans le rôle des déflections et vitesses
de débattement maximales : elles sont désormais vues comme des contraintes de la discipline
Qualités de Vol Boucle Fermée, et non plus comme des résultats.
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
216 Commande
B.2 Obtention d’un modèle aérodynamique paramétré

B.2.1 Paramétrisation géométrique des gouvernes de bord de fuite

Les procédures “classiques” de dimensionnement conjoint de lois de commande et systèmes phy-


siques présentées dans la littérature s’attachent principalement à minimiser la surface du plan
horizontal sous contraintes longitudinales. L’aile volante étant par définition dépourvue de plan
horizontal, un tel processus est inapplicable dans notre cas. C’est pourquoi l’envergure totale des
élevons du bord de fuite est choisie comme variable à optimiser. Plus précisément une variable
η représentant le rapport de l’envergure des élevons externes sur leur envergure initiale est in-
troduite. Sur la figure B.3(a) l’architecture initiale, correspondant à η = 1, est représentée. Cette
configuration initiale est composée de 5 gouvernes de chaque côté de l’aile et s’étalant sur tout
le bord de fuite, hormis un espace entre l’élevon 1 et 2 pour le mât moteur.

A partir de cette configuration initiale, on définit la variable η avec les hypothèses suivantes :

y2inb
• η = y2init
, y2inb et y2init
inb
étant respectivement les positions internes en y de l’élevon 2
inb
paramétré et initial respectivement.

• y5out = y5init
out
, y5out et y5init
out
étant respectivement les positions internes en y de l’élevon 5
paramétré et initial respectivement.

• Les cordes relatives des élevons sont constantes et égales à 22%.

• Les élevons 2 à 5 sont découpés de manière égale.

• L’élevon 1 reste constant.

Il apparaît alors clairement que 0 ≤ η ≤ 1, η = 1 correspondant à la configuration initiale


et η = 0 correspondant à la configuration où il n’y a plus aucun élevon sur l’aile externe. Des
exemples d’architecture pour η = 0.8 et η = 0.4 sont représentées aux figures B.3(b) et B.3(c)
respectivement.

(a) η = 1 (b) η = 0.8 (c) η = 0.4

Figure B.3 : Taille des gouvernes externes pour diférentes valeurs du paramètre η.
B.2. Obtention d’un modèle aérodynamique paramétré 217

B.2.2 Calcul de modèles aérodynamiques pour des envergures de gouvernes dis-


crétisées

Dans cette section nous décrirons le calcul de modèles aérodynamiques pour différentes valeurs
de η . Pour des raisons de clarté l’architecture initiale – c’est-à-dire la configuration η = 1 –
est appelée “avion référence”, tandis que les configurations modifiées – c’est-à-dire pour η <1
– sont appelées avions “projet”. La forme en plan ainsi que les profils sont constants sur toutes
les configurations, c’est pourquoi seules les efficacités aérodynamiques des gouvernes doivent
être calculées sur l’avion projet. Encore une fois l’objectif de cette section est d’obtenir une
paramétrisation continue par η de la représentation d’état de l’avion. La procédure permettant
d’obtenir une telle approximation continue est représentée en figure ??. Une première étape est
de calculer les modèles aérodynamiques calibrés pour plusieurs valeurs de η , plus précisément
pour des valeurs allant de 0.1 à 1, par pas de 0.1. A cet effet nous avons utilisé le logiciel AVL ,
associé à des facteurs de calibration provenant de la configuration initiale que l’on suppose connue
d’un point de vue aérodynamique. Ceci permet de prendre en compte des effets de Mach et des
non-linéarités à forte incidence et forte déflection, ce qu’AVL ne peut pas prédire.

Plus précisément le coefficient de portance globale avion Cz comprend une contribution de


la déflection de la i-ème gouverne Cz i qui peut être écrite sous la forme suivante :

NL
Cz i = kδmi
(α, δmi )Czδmi δmi (B.1)

où kδmN L (α, δm ) est un facteur représentant la perte d’efficacité à forte incidence et forte dé-
i i
N L (α, δm )
flection, et Czδmi est le gradient de portance lié à la déflection de la i-ème gouverne. kδm i i
est connu pour l’avion de référence, et on le conserve sur l’avion projet. Le gradient de portance
de référence Czδmi ref est supposé connu lui aussi d’études précédentes [Meheut et al., 2012] et
est comparé avec le gradient de portance calculé par AVL pour la configuration de référence
Czδmi AV L (voir la figure B.5(a)). A partir de ces données un facteur de calibration permettant
de tenir compte du manque de précision d’AVL est calculé :

Czδmi ref
∆Cz AV
i
L
= (B.2)
Czδmi AV L

Ensuite un calcul AVL est lancé sur l’avion projet afin de calculé le gradient de portance
proj
Czδmi AV L . Ce gradient est finalement calibré avec le facteur de calibration ∆Cz AV
i
L , ce qui

donne :

proj
Cz proj
i
NL
= kδm i
∆Cz AV
i
L
Czδmi AV L δmi (B.3)

Une procédure similaire est appliquée afin de calculer les moments de tangage pour la configu-
ration projet. Plus précisément le moment de tangage global avion Cm possède une contribution
du braquage de la i-ème gouverne Cmi qui peut s’écrire :
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
218 Commande

NL
Cmi = kδmi
(α, δmi )Czδmi δmi (XCG − XFi ) (B.4)

où XCG est la position en x du centre de gravité, et XFi est la position en x du foyer de


la gouverne i. Cependant XFi n’est pas directement une sortie d’AVL, et nécessite d’être calculé
via la formule suivante :

AV L
Cm
AV L δm i
XFi = XCG − (B.5)
CzAV
δm
L
i

Ceci permet de calculer le facteur de calibration du foyer aérodynamique à partir de la connais-


sance du foyer de la gouverne i XFi ref :

XFi ref
∆XFi AV L = (B.6)
XFi AV L

proj
Finalement on calcule via AVL le moment de tangage de l’avion projet CmG
δmi
AV L , et le
foyer aérodynamique résultant est calculé et calibré comme suit :

proj
 
AV L
Cm
proj δm
XFi AV L = ∆XFi AV L . XCG − Lproj
i  (B.7)
CzAV
δm i

Un procédé similaire sert à déterminer les efficacités en roulis des gouvernes pour l’avion projet.
Plus précisément le coefficient de roulis de l’avion Cl comprend une contribution du braquage
de la gouverne i Cli qui peut être écrit de la façon suivante :

NL
Cli = kδmi
(α, δmi )Clδmi δmi (B.8)

Le facteur de calibration pour le coefficient de moment en roulis est alors déterminé par :

Clδmi ref
∆ClAV
i
L
= (B.9)
Clδmi AV L

Et finalement le coefficient de roulis de l’avion projet est calibré grâce au facteur de calibration
précédemment calculé :

proj
Clproj
i
NL
= kδmi
∆ClAV
i
L
Clδmi AV L δmi (B.10)

Cette méthode, schématisée en figure B.4, regroupe les avantages d’une génération rapide
B.2. Obtention d’un modèle aérodynamique paramétré 219

de résultats grâce à l’utilisation de code CFD linéaire, et une précision accrue par rapport aux
résultats bruts d’AVL grâce à la connaissance de la configuration de référence. Ici ce procédé est
appliqué uniquement aux coefficients aérodynamiques liés aux gouvernes, mais il serait similaire
pour le calcul de n’importe quel coefficient d’un avion projet à partir de connaissance d’un avion
de référence.

Avion de Référence
Modèle Aérodynamique

Avion avec AVL Bibliothèque


Taille des Elevons variant Calcul Modèles Aérodynamiques ABRICOT
Modèle Géométrique Aérodynamique Paramétrés Approximation LFR
η = 0.1 : 1
η = 0.1 : 1
masse Mach H masse Mach H

LFR fonction de
l’envergure élevons

masse Mach H

Figure B.4 : Processus d’obtention de représentation d’état paramétrées sous formalisme LFR,
avec comme paramètre variant η.

B.2.3 Approximation polynômiale et représentation LFR des efficacités aérody-


namiques de gouvernes

Une fois que les coefficients aérodynamiques sont calculés et calibrés comme décrit précédemment,
l’étape finale consiste à d’obtenir une approximation continue de représentation d’état paramétrée
par η . Pour cela nous avons travaillé avec le cadre formel des LFR , car cette représentation
est adaptée à l’algorithme d’optimisation pour lois de commande utilisé par la suite. De plus de
puissants algorithmes permettant de convertir un jeu discret de données en une représentation
LFR continue ont été développés par l’Onera [Roos et al., 2014]. Une LFR est un modèle dans
lequel tous les paramètres invariants sont regroupés au sein d’un unique bloc M , tandis que les
incertitudes et paramètres variants sont contenus au sein d’une matrice bloc-diagonale ∆. Les
fonctions polynômiales et rationnelles sont par exemple convertibles en LFR .

Plus précisément, le problème est de trouver une LFR approximant le plus précisément possible
les différentes représentations d’état calculées pour différentes valeurs de η . Les incertitudes
ne sont pas considérées dans notre étude, c’est pourquoi le bloc ∆ est composé uniquement du
paramètre η répété plusieurs fois. De plus, ainsi que mentionné précédemment, dans notre travail
la forme en plan de l’aile volante reste constante, c’est pourquoi seule la matrice de commande
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
220 Commande
0.5

0.4 0.15

0.3
Czδmi (1/rad)

Czδmi (1/rad)
0.10
0.2

0.1 0.05

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Mach Mach
elevon 1 2 3 4 5 elevon 1 2 3 4 5
(a) Comparaison des gradients de portance pour (b) Comparaison des gradients de portance de l’avion
l’avion de référence : CLref
δmi
(dotted) vs CLAV L
δmi
projet avant et après calibration respectivement :
Lproj
(plain). CLAV
δmi
(plain) vs CLproj
δmi
(dotted). L’avion projet
étudié ici est η = 0.6.

Figure B.5 : Comparaison des gradients de portance CLδmi en sortie d’AVL avec les données de
référence (gauche) et les données calibrées (droite).

(a) CmLF R
δmi pour η variant.
LF R
(b) Approximation polynômiale de Clδmi
pour η
variant.

Figure B.6 : Approximations polynômiales LFR des gradients de tangage et roulis.

B regroupant les efficacités des élevons est approximée par la LFR . Cependant la résolution du
problème serait identique en cas d’inclusion de variables ayant un effet sur la matrice A .

Nous avons de plus décidé de se restreindre à des approximations polynômiales afin de conserver
l’ordre de la LFR – c’est-à-dire le nombre de fois où η est répété au sein du bloc ∆ aussi petit
que possible. D’un point de vue physique ceci peut être justifié par le fait que les efficacités de
gouverne varient de manière continue et lisse par rapport à leur envergure. Nous avons utilisé
la routine lsapprox de la bibliothèque APRICOT sur Matlab [Roos et al., 2014]. Le problème
consiste à trouver un polynôme P de degré np minimisant le critère suivant :

N
X
C= [Bηk − P (η)]2 (B.11)
k=1
B.3. Codesign de gouvernes et lois de commandes sous contraintes qualités de vol 221

Bηk étant la matrice de commande pour une valeur discrétisée de η , η k,k=1:N = [0.1...1].

Le degré du polynôme P est défini à np = 5, et la LFR résultante est de taille 20 – ceci repré-
sente le nombre de fois où η est répété au sein du bloc ∆. L’erreur RMS maximale vaut 9.36.10−3
et l’erreur maximale absolue vaut 2.01.10−2 . Les approximations résultantes sont représentées en
figure B.6.

B.3 Codesign de gouvernes et lois de commandes sous contraintes


qualités de vol

Dans cette section nous développons le problème de dimensionnement conjoint qui consiste à
simultanément minimiser le paramètre η tout en satisfaisant des contraintes de qualités de vol
et de manœuvrabilité.

B.3.1 Structure des lois de commande

Comme déjà précisé plus haut, l’instabilité longitudinale sur cette aile volante requiert un système
de stabilisation pour rendre l’avion sûr. De plus des lois de commande latérales sont également
nécessaires pour améliorer les qualités de vol latérales. La prise en compte de ces lois de commande
à la fois longitudinale et latérale est de plus nécessaire pour un dimensionnement adéquat des
gouvernes. Une contribution importante de notre travail est alors de fournir une méthodologie
pour une synthèse simultanée des lois sur les trois axes, avec une structure de lois arbitraire, alors
que ce problème est habituellement traité en découplant l’axe de tangage des axes de roulis / lacet.
Dans cette étude nous considérons une architecture typique de commandes de vol électriques. Le
pilote fournit des entrées en termes de facteur de charge commandé nzc , inclinaison φc et
dérapage βc . La loi de commande stabilisante est appelée C∗ et Y ∗ en longitudinal et latéral
respectivement, avec une structure décrite dans [Favre, 1994].

Un aperçu général de la structure de lois est fourni en figure B.7. Ainsi que déjà mentionné, les
paramètres de sortie de la loi sont des ordres équivalent de braquage profondeur, aileron et direction
respectivement, qui sont indépendants de l’architecture de gouvernes. Ces ordres équivalents sont
alors convertis en braquage de gouvernes grâce à un module d’allocation décrit dans la section
suivante.

B.3.2 Modèle d’allocation

L’allocation de commandes est le problème de convertir des ordres équivalents calculés par la loi
de commande en braquages de gouvernes quand il y a plus de gouvernes que d’axes à contrôler.
Dans notre étude il est nécessaire d’incorporer un module d’allocation permettant de convertir des
ordres équivalents profondeur, aileron et direction (δmequi , δlequi , δnequi ) en véritable braquage de
gouvernes δmi,i=1..10 , δn.
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
222 Commande
Mathématiquement, le problème d’allocation consiste à trouver un vecteur u de déflections
satisfaisant :    
Cmδm1 ... Cmδm10 Cmδn Cmδmequi δmequi
 Clδm1 ... Clδm10 Clδn  u =  Clδlequi δlequi  (B.12)
   

Cnδm1 ... Cnδn10 Cnδn Cnδnequi δnequi


| {z }
B1 (η)

où B1 (η) est la matrice des efficacités de gouvernes en tangage, roulis, et lacet respectivement,
tous dépendants du paramètre η . [Cmδmequi Clδlequi Cnδnequi ]T est un vecteur de gradients
équivalents vus par la loi de commande. Ces valeurs peuvent être choisies arbitrairement sans
perte de généralité, dans notre étude nous choisissons une valeur de 1 sur chaque axe. Alors une
solution classique de l’équation précédente est la pseudo-inverse de Moore-Penrose :
 
δmequi
u = Kalloc (η)  δlequi  (B.13)
 

δnequi
avec Kalloc (η) = B1T (B1 B1T )−1 (B.14)

Ainsi que discuté en détail par la suite, notre procédure comporte 2 étapes distinctes. Dans
une première étape, les gains de lois sur les 3 axes sont calculés pour une valeur fixe de η ; lors de
cette étape l’allocation décrite dans l’équation précédente est utilisée. Lors de la 2ème étape, les
gains des lois 3 axes ainsi que le paramètre η sont optimisés conjointement. Dans cette étape ;
Kalloc n’est plus fixe, mais une variable d’optimisation. Grâce à cela l’espace de solutions est
agrandi, et une stratégie véritablement optimale peut être trouvée par l’algorithme d’optimisation.

Plus précisément, afin de limiter le nombre de variables pour l’optimisation, Kalloc est
paramétré de la façon suivante :
pitch

roll

Kalloc Kalloc 0
 pitch roll
Kalloc = Kalloc −Kalloc 0  (B.15)

yaw
0 0 Kalloc

Ceci signifie que nous imposons un braquage symétrique et anti-symétrique pour un ordre en
tangage et roulis respectivement, et un ordre en lacet sera alloué à la gouverne de direction uni-
quement. En incorporant cette connaissance physique dans la structure de l’allocation, le nombre
de variables au sein de Kalloc est réduit de 33 à 11.

B.3.3 Synthèse simultanée de lois 3 axes

La synthèse conjointe appliquée à l’aile volante décrite ici suit un schéma en 2 temps :

1. Une première synthèse de lois calcule les gains pour une valeur de η aribtaire et fixée.
Le problème consiste en la minimisation de la différence entre un modèle de référence et
B.3. Codesign de gouvernes et lois de commandes sous contraintes qualités de vol 223

ω02 nref
z + z̃1
s2 +2ξ0 ω0 s+ω02 −
Dynamique Longi Référence
1 φref + z2
(1+s/τrp )(1+s/τsp )

Dynamique Roulis Référence
2
ωdr βref + z3
2
s2 +2ξdr ωdr s+ωdr

Dynamique Retard Référence

η η α Ve
g zα
nz
q
Allocation
δmequi Représentation LFR β
nzc Loi C ∗ Pseudo-Inverse δmi i = 1..10 2
ωact u Avion Retard
LFR δn s2 +2ξact ωact s+ωact2 p
φc δlequi Modèle Actionneurs r
βc Loi Y ∗ δnequi ew q 1+ V

3Lz
s wz φ
σz 2L z e
πVe (1+ Lz s)2
Ve

Modèle Turbulence

Figure B.7 : Schéma de principe du codesign avec allocation paramétrée. Les blocs tunables
sont présentés en orange.

l’avion en boucle fermée. La sortie de cette étape est la valeur optimale du critère H∞ .
Cette section est dédiée à la description de cette étape.

2. La sortie de l’étape 1 est utilisée comme contrainte sur la valeur maximale du critère H∞ ,
afin de garantir un comportement satisfaisant en boucle fermée tout en optimisant la taille
des élevons η . Cette étape est décrite en détail dans la section suivante.

Ainsi que déjà précisé, l’optimisation de la taille gouvernes η et des lois de commande requiert
d’imposer une contrainte – au sens de l’optimisation – qui garantit un comportement adéquat en
boucle fermée de la solution optimale. Cette contrainte est imposée sur la valeur maximale d’une
valeur admissible d’un critère H∞ , qui doit être calculée lors d’une première synthèse. Le critère
H∞ et le calcul de sa valeur optimale sont décrits ici.

Nous avons utilisé un schéma de suivi de modèle de référence trois-axes. Ce schéma consiste en
la minimisation de la différence entre un modèle de dynamique de référence et l’avion en boucle
fermée, du point de vue de la norme H∞ . Si le modèle de référence est parfaitement recopié
par la boucle fermée sur l’ensemble du domaine de vol, alors la valeur du critère H∞ optimal
est 0. Cependant ceci est impossible à réaliser d’un point de vue pratique à cause de limitations
physiques, c’est pourquoi la valeur optimale est toujours au-delà de 0.

Ici le modèle de référence est écrit comme une minimisation de la norme d’un transfert 3 × 3
entre les entrées pilote (nzc , φc , βc et les sorties (z1 , z2 , z3 ) les différences entre la dynamique de
référence et la dynamique boucle fermée (nz , φ, β). Le lecteur peut se référer à la figure B.7 pour
une définition exacte des signaux. Dans notre cas, la spécification par transferts multi-canaux a
plusieurs avantages par rapport à de multiples transferts simples. Tout d’abord les termes hors
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
224 Commande
diagonale sont implicitement optimisé à 0. Ainsi la loi de commande résultante va totalement
découpler les 3 axes, c’est-à-dire le longitudinal du latéral, mais également le roulis du lacet – les
virages se font sans dérapage – et inversement – un ordre palonnier ne provoque pas d’inclinaison.
Les couplages entre tous les axes sont explicitement pris en compte par la loi de commande, ce
qu’une approche SISO ne permettrait pas. Enfin cette formulation est indépendante du modèle
boucle ouverte et de la structure de lois de commande. Enfin cela permet une formulation élégante
du problème d’optimisation de la section suivante : une contrainte unique sur la norme H∞ de ce
transfert multi-canaux permet un comportement adéquat en boucle fermée sur les trois axes. Une
formulation équivalente avec des transferts SISO requerrait neuf contraintes.

Les modèles de référence selon les axes longitudinal, latéral et directionnel sont définis respec-
tivement comme suit :

N zref ωi2
= (B.16)
N zc s2 + 2ξi ωi s + ωi2
φref 1
= (B.17)
φc (1 + τrp s)(1 + τsp s)
βref 2
ωdr
= 2 (B.18)
βc s2 + 2ξdr ωdr s + ωdr

Dans notre étude, les valeurs pour la boucle fermée de référence sont données dans la table
suivante. Idéalement ces valeurs devraient dépendre de la dynamique en boucle ouverte, ceci sera
l’objet d’une étude future.

Parameter ωi ξi τrp τsp ωdr ξdr


Value 1 rad.s−1 0.7 1.6 rad.s−1 1.9 rad.s−1 0.4 rad.s−1 0.7

Table B.1 : Valeurs pour le modèle de référence.

Le problème d’optimisation pour l’optimisation simultanée des lois 3 axes s’écrit donc simple-
ment :

min T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P (η), K) (B.19)


K ∞
tel que :
∀p ∈ C, p pôle de P (s) :
Re(p) ≤ −MinDecay, Re(p) ≤ −MinDamping.|p|
Kstabilise P (η)

K étant un vecteur contenant tous les gains des lois définis dans la section précédente. Les
contraintes additionnelles s’assurent que les pôles en boucle fermée ont un amortissement d’au
B.3. Codesign de gouvernes et lois de commandes sous contraintes qualités de vol 225

moins 0.5, et une partie réelle d’au moins -0.2. Bien que ces contraintes puissent à première vue
être redondantes avec les objectifs du suivi de modèle de référence, nous avons remarqué qu’elles
facilitaient la convergence de l’algorithme d’optimisation.

Pour résoudre ce problème d’optimisation, on a utilisé la routine systune de la Robust Control


Toolbox de Matlab. Cette fonction permet l’optimisation de contrôleurs d’ordre fixe, ainsi elle peut
prendre en compte des paramètres physiques en plus du calcul des gains. Le couplage entre le
problème de dimensionnement et de contrôle peut donc être résolu en une seule optimisation. De
plus cette fonction est adaptée aux propriétés mathématiques du problème d’optimisation décrit
dans l’équation B.19, à savoir le fait que la norme H∞ soit non-lisse. Les optimums trouvés
par l’algorithme sont donc locaux, par conséquent on veillera à répéter le calcul avec différentes
initialisations afin de s’assurer de la globalité de la solution.

Un exemple de lois de commande multi-axes après optimisation est présenté en figures B.8 et
B.9 à travers les réponses fréquentielles et temporelles respectivement. Le point de vol présenté
est à M .35, H = 3300f t, m = 300T . La valeur optimale du critère H∞ pour ce point de vol
γ∞ , ainsi que la valeur maximale des contraintes normalisées à l’optimum gBest , sont :

γ∞ = 0.107 (B.20)
gBest = 0.9997

gBest < 1 à l’optimum signifie que les contraintes sont satisfaites.

Afin de visualiser le comportement en boucle fermée de la loi de commande résultante, la


simulation présentée en figure B.9 comprend les manœuvres suivantes :

• à t = 0s, une manœuvre à cabrer est commandée via un échelon ∆nzc = 1g. Par conséquent
les braquages des gouvernes gauche et droite sont symmétriques.

• à t = 5s, un échelon d’inclinaison φc = 30˚est commandé. Les élevons sont alors braqués de
manière anti-symmétrique. La gouverne de direction est également activée afin de conserver
un dérapage nul.

• à t = 12s, un dérapage de βc = 5˚ est commandé.

Les braquages initiaux permettant d’équilibrer l’avion ne sont pas représentés sur la figure, unique-
ment les variations autour de leur position d’équilibre. En conclusion de cette partie, nous insistons
sur le fait que le résultat attendu de cette partie est uniquement la valeur optimale γ∞ du critère
H∞ , qui sera utilisé comme contrainte du problème d’optimisation décrit dans la section suivante.
En particulier les gains de la loi de commande ne sont pas conservés, ils sont à nouveau calculés
lors du codesign de η et des lois de commande.
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
226 Commande

nref
z /nzc φref /φc
nz /nzc φ/φc
z1 /nzc z2 /φc

(a) [Comparaison des réponses fréquentielles du modèle (b) [Comparaison des réponses fréquentielles du modèle
de référence longitudinal, de la boucle fermée et de l’er- de référence latéral, de la boucle fermée et de l’erreur
reur après minimisation de T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } ∞ , après minimisation de T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } ∞ , avec
avec contrainte sur le lieu des pôles. contrainte sur le lieu des pôles.

βref /βc
β/βc
z3 /βc

(c) [Comparaison des réponses fréquentielles du modèle


de référence en lacet, de la boucle fermée et de l’erreur
après minimisation de T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } ∞ , avec
contrainte sur le lieu des pôles.

Figure B.8 : Diagrammes de Bode en amplitude selon les 3 axes des modèles de référence et
boucle fermée résultant de la minimisation de T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } avec contrainte sur le lieu

des pôles.

B.3.4 Définition du problème d’optimisation

Une fois que l’on a calculé l’optimum γ∞ du problème d’optimisation, on l’utilise comme
contrainte sur la norme H∞ du transfert multi-canaux du problème conjointe d’optimisation.
Plus précisément, le problème conjoint de dimensionnement et contrôle consiste à trouver une
valeur minimale de η telle que :
B.3. Codesign de gouvernes et lois de commandes sous contraintes qualités de vol 227

Control Surfaces Response


Elevons Deflections (in ˚)

δm1
δm2
δm3
δm4
δm5
δn
Time (s)
Control Surfaces Deflection Rates
Eflection Rates ((in ˚/s)

˙ 1
δm
˙ 2
δm
˙ 3
δm
˙ 4
δm
˙ 5
δm
˙
δn

Time (s)
Aircraft Response

q
q,β,p,r,φ

β
p
nz

r
φ
nz

Time (s)

Figure B.9 : Réponses temporelles pour η = 1 (pointillés) et η opt = 0.3885 (traits pleins).
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
228 Commande
• Les braquages et vitesses de braquage en réponse à une manœuvre ne dépassent pas certaines
limites.

• Le transfert boucle fermée est optimal.

On s’assure du dernier point en contraignant la norme H∞ du transfert multi-entrées multi-


sorties T{nzc ,φc ,βc }→{z̃1 ,z2 ,z3 } (P, K) à ne pas dépasser la valeur optimale γ∞ – c’est-à-dire

à rester égale à cette valeur – quel que soit η . Une fois la valeur γ∞ connue pour une valeur
initiale de η grâce au procédé décrit dans la section précédente, le problème d’optimisation
consistant à trouver le meilleur comportement en boucle fermée se traduite en un unique problème
de satisfaction de contraintes. Cette formulation élégante permet de minimiser un autre objectif
– η en l’occurrence – tout en garantissant un comportement satisfaisant en boucle fermée sur
tous les axes.

Les contraintes de manœuvrabilité sur chaque axe sont exprimées en tant que contraintes
H∞ de fonctions de transfert appropriées, tel que décrit dans les études de [Niewoehner and
Kaminer, 1996; Kaminer et al., 1997]. C’est pourquoi ces contraintes sont sur l’écart-type et non
pas temporelles.

On garantit les contraintes suivantes :

• Braquage et vitesse de débattement maximal à une action longitudinale du pilote. On s’as-


sure de cette contrainte en contraignant les normes H∞ suivantes : ∆δm1i max Tnzc →δmi ∆nzc max ≤

1 max
1 et ˙ i max Tnzc →δm
δm ˙ i ∆nzc ≤ 1 respectivement, avec ∆δmi max = δmi max − δmie .

• Braquage et vitesse de débattement maximal à une action en roulis du pilote. On s’assure de


cette contrainte en contraignant les normes H∞ suivantes : ∆δm1i max Tφc →δmi φc max ≤1

1 max
et ˙ max Tφc →δm
δm ˙ i φc ≤1
i ∞

• Braquage et vitesse de débattement maximal nécessaire à stabiliser l’avion dans une turbu-
lence sévère longitudinale. On s’assure de cela en contraignant les normes H∞ suivantes :
2
∆δmi max Tew →δmi ≤ 1 et δm˙ 2max Tew →δm˙ i ≤ 1.
∞ i ∞

On choisit les valeurs suivantes pour le dimensionnement :

• δmi max = 25˚. On a conservé une marge de 5˚par rapport à la butée mécanique de 30˚, afin
d’éviter de se trouver dans la partie à efficacité réduite des gouvernes. Notons également
que le braquage maximale autorisé tient compte du braquage nécessaire à l’équilibrage δmie
. Plus on a besoin de braquage pour équilibrer l’avion, moins il reste de braquage pour le
contrôle.
˙ i max = 60˚/s.
• δm

• ∆nzc max = 1.5g.

• φc max = 45˚.
B.4. Résultats 229

B.4 Résultats

B.4.1 Codesign sur un seul point de vol

Le codesign avec allocation tunable appliqué au point de vol M .35, H = 3300f t, m = 300T ,
donne les résultats suivants :

η opt = 0.3885 (B.21)


gBest = 0.998 (B.22)

On parvient à diminuer de 61% l’envergure des gouvernes par rapport à la configuration initiale,
tout en satisfaisant les contraintes de qualités de vol – c’est-à-dire gBest < 1. Plus précisément
la figure B.10 rassemblant les contraintes qualités de vol à l’optimum, indique qu’en ce point
de vol c’est la contrainte de roulis qui est limitante. Par conséquent, l’allocation optimale est de
distribuer de manière égale sur toutes les gouvernes l’ordre en roulis, afin de saturer de manière
égale toutes les contraintes de roulis. Le tangage n’étant pas limitant sur ce point de vol, seulement
deux gouvernes sont utilisées pour le contrôle en tangage. Cette stratégie est également visible
sur la réponse temporelle de la figure B.9 : les cinq gouvernes sont utilisées de manière égale
pour la manœuvre en roulis, se traduisant par des vitesses de débattement égales pour toutes les
gouvernes. Ce comportement est confirmé par la matrice d’allocation :
 
0.023 0.3811 0
0.0116 0.3811 0 
 
 
 0.000 0.3811 0 
 
 0.000 0.3811 0 
 
 0.000 0.3811 0 
 
opt  
Kalloc =
 0.023 −0.3811 0  (B.23)
0.0116

−0.3811 0 
−0.3811
 
 0.000 0 
 
 0.000
 −0.3811 0 
 0.000 −0.3811 0 
 

0.000 0.000 0.006

B.4.2 Résultats sur l’ensemble du domaine de vol

On calcule ensuite le codesign avec allocation tunable sur l’ensemble du domaine de vol. La
figure B.11 représente la surface résultante η opt = f (M , H), ainsi que les contraintes limitantes
associées à chaque optimum. La valeur maximale est η opt = 0.5996 : l’avion peut être contrôlé de
manière adéquate sur l’ensemble du domaine de vol, sans atteindre ses butées, avec uniquement
60% de l’envergure initiale de gouvernes. De plus les valeurs dimensionnantes de η opt sont
atteintes pour des cas de vol à faible pression dynamique, c’est-à-dire à haute altitude et basse
Annexe B. Dimensionnement Préliminaire de Surfaces de Contrôle et Lois de
230 Commande
1

0.9

0.8

0.7

0.6
H∞ norms of constraints

0.5

0.4

0.3

0.2

0.1

0
δm1 δm2 δm3 δm4 δm5
Elevon number

1 max (P (η opt , K opt ))


∆δmi max Tnzc →δmi ∆nzc ∞
1 max (P (η opt , K opt ))
T
˙ i max nzc →δm
δm ˙i ∆n zc

2 opt , K opt ))
∆δmi max T ew →δm i
(P (η

2 opt , K opt ))
˙
δmi max T ˙
ew →δmi (P (η

1 max (P (η opt , K opt ))
T
∆δmi max φc →δmi c φ

1 max (P (η opt , K opt ))
T
˙ i max φc →δmi
δm ˙ φ c

Figure B.10 : Diagramme en barres des contraintes normalisées après optimisation, η opt =
0.3885.

vitesse. Très intéressant également, ces points sont dimensionnés à la fois par de la turbulence,
de la manœuvre longitudinale et latérale. Comme prévu, la stabilisation à une entrée exogène est
vraiment dimensionnante pour certains points de vol. De plus ces résultats prouvent l’importance
de l’allocation dans le dimensionnement des gouvernes : l’allocation a une influence directe sur le
dimensionnement.
B.4. Résultats 231

Figure B.11 : η opt en fonction du cas de vol, calculé via le codesign avec allocation tunable.

Conclusion et Perspectives

Nous avons présenté ici une nouvelle méthode de dimensionnement pour une aile volante instable.
Cette méthode consiste à optimiser simultanément les lois de commande longitudinale, latérales
ainsi qu’un module d’allocation, tout en minimisant l’aire des gouvernes sous contraintes qualités
de vol. D’un point de vue dimensionnement, nous avons montré que l’aile volante Airbus peut
être contrôlée de manière adéquate sur tous les axes, sur l’ensemble du domaine de vol, avec
uniquement 60% de l’envergure initiale de gouvernes, en loi normale et pleine autorité de gouverne,
avec des hypothèses raisonnables sur les actionneurs, capteurs et chaîne d’acquisition. Des travaux
futurs incluront des cas de panne.

D’un point de vue méthodologique, nous avons démontré l’intérêt d’une méthode de dimen-
sionnement couplée par rapport à une approche itérative : en une seule étape le dimensionnement
garantit les contraintes qualités de vol, tout en tenant compte des lois de commande. Une perspec-
tive de travail concerne l’inclusion de critères temporels plutôt que fréquentiels. Enfin on pourra
tester d’autres optimiseurs, par exemple des routines basées sur de la simulation telles que des
algorithmes génétiques.
Bibliography
Abzug, M. J. and Larrabee, E. E. (2005). Airplane stability and control: a history of the tech-
nologies that made aviation possible, volume 14. Cambridge University Press.

Alazard, D. (2013). Reverse Engineering in Control Design. John Wiley & Sons.

Alazard, D., Cumer, C., Apkarian, P., Gauvrit, M., and Ferreres, G. (1999). Robustesse et
commande optimale. Cépaduès-éditions.

Alazard, D., Loquen, T., de Plinval, H., and Cumer, C. (2013). Avionics/Control co-design for
large flexible space structures. In AIAA Guidance, Navigation, and Control (GNC) Confer-
ence, Guidance, Navigation, and Control and Co-located Conferences. American Institute of
Aeronautics and Astronautics.

Anderson, C. W. (1989). Learning to control an inverted pendulum using neural networks. Control
Systems Magazine, IEEE, 9(3):31–37.

Anderson, M. and Mason, W. (1996). An MDO approach to Control-Configured-Vehicle Design.


In AIAA, volume 4058, Bellevue, WA. AIAA.

Apkarian, P. (2012). Tuning Controllers Against Multiple Design Requirements. In System Theory,
Control and Computing (ICSTCC), 2012 16th International Conference on, pages 1–6. IEEE.

Apkarian, P. and Noll, D. (2006). Nonsmooth H∞ synthesis. Automatic Control, IEEE Transac-
tions on, 51(1):71–86.

Apkarian, P., Noll, D., Rondepierre, A., and others (2007). Nonsmooth optimization algorithm
for mixed H2/H∞ synthesis. In Proc. of the 46th IEEE Conference on Decision and Control,
pages 4110–4115.

Ashkenas, I. L. and Klyde, D. H. (1989). Tailless aircraft performance improvements with relaxed
static stability.

Aström, K. J. and Kumar, P. (2014). Control: A perspective. Automatica, 50(1):3–43.

Bansal, V., Ross, R., Perkins, J., and Pistikopoulos, E. (2000). The interactions of design and
control: double-effect distillation. Journal of Process Control, 10(2–3):219–227.

Beaverstock, C. S., Richardson, T. S., Maheri, A., and Isikveren, A. (2012). Integrated flight con-
trol system development using CEASIOM. Canadian Aeronautics and Space Journal, 58(1):29–
46.

Belschner, T. (2011). Definition of systems architecture for flight, load and vibration control of
blended wing body type aircraft. MSc Thesis-Airbus Internal D3-21, EADS Innovation Works.

Bérard, C., Biannic, J.-M., and Saussié, D. (2012). La commande multivariable: Application au
pilotage d’un avion. Dunod.

233
234 Bibliography

Bodson, M. (2002). Evaluation of Optimization Methods for Control Allocation. JOURNAL OF


GUIDANCE, CONTROL, AND DYNAMICS, 25(4):703–711.

Bodson, M. and Frost, S. A. (2011). Load Balancing in Control Allocation. Journal of Guidance,
Control, and Dynamics, 34(2):380–387.

Boiffier, J.-L. (1998). The dynamics of flight: the equations, volume 1. John Wiley & Sons.

Bolsunovsky, A., Buzoverya, N., Gurevich, B., Denisov, V., Dunaevsky, A., Shkadov, L., Sonin, O.,
Udzhuhu, A., and Zhurihin, J. (2001). Flying wing—problems and decisions. Aircraft Design,
4(4):193–219.

Bordignon, K. and Bessolo, J. (2002). Control Allocation for the X-35b. In 2002 Biennial Interna-
tional Powered Lift Conference and Exhibit, Aviation Technology, Integration, and Operations
(ATIO) Conferences. American Institute of Aeronautics and Astronautics.

Bradley, K. R. (2004). A sizing methodology for the conceptual design of blended-wing-body


transports. NASA CR, 213016:2004.

Brinker, J. S. and Wise, K. A. (1996). Stability and flying qualities robustness of a dynamic
inversion aircraft control law. Journal of Guidance, Control, and Dynamics, 19(6):1270–1277.

Brinker, J. S. and Wise, K. A. (2001). Flight Testing of Reconfigurable Control Law on the X-36
Tailless Aircraft. JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS, 24(5).

Buffington, J. (1997). Tailless Aircraft Control Allocation. Technical report, Wright Laboratory-
Air Force Materiel Command.

Burke, J., Henrion, D., Lewis, A., and Overton, M. (2006). HIFOO -a MATLAB package for
fixed-order controller design and H∞ optimization.

Burken, J. (2001). Two Reconfigurable Flight-Control Design Methods: Robust Servomechanism


and Control Allocation. JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS, 24(3):482–
493.

Calderara, F. and Lemaignan, B. (2007). Stabilité longitudinale- Rapport technique misco. Airbus
internal RP0709670, Toulouse, France.

Cameron, D. and Princen, N. (2000). Control allocation challenges and requirements for the
Blended Wing Body. In AIAA Guidance, Navigation, and Control Conference and Exhibit,
Guidance, Navigation, and Control and Co-located Conferences. American Institute of Aero-
nautics and Astronautics.

Chakraborty, I., Jackson, D., Trawick, D. R., and Mavris, D. (2013a). Development of a Sizing
and Analysis Tool for Electrohydrostatic and Electromechanical Actuators for the More Electric
Aircraft. American Institute of Aeronautics and Astronautics.

Chakraborty, I., Trawick, D. R., Jackson, D., and Mavris, D. (2013b). Electric Control Surface
Actuator Design Optimization and Allocation for the More Electric Aircraft. In 2013 Aviation
Technology, Integration, and Operations Conference, Los Angeles, CA. AIAA.
Bibliography 235

Chalk, C. (1969). MIL-F-87 military specifications-flying qualities of piloted aircrafts.

Chambers, J. (2005). INNOVATION IN FLIGHT: RESEARCH OF THE NASA LANGLEY RE-


SEARCH CENTER ON REVOLUTIONARY ADVANCED CONCEPTS FOR AERONAUTICS.

Chudoba, B. (2001). Development of a generic stability and control methodology for the concep-
tual design of conventional and unconventional aircraft configurations. PhD thesis, Cranfield
University.

Coleman, G. and Chudoba, B. (2007). A Generic Stability and Control Tool for Conceptual Design,
Prototype Styem Overview. In 45th AIAA Aerospace Sciences Meeting and Exhibit, Aerospace
Sciences Meetings. American Institute of Aeronautics and Astronautics.

De Castro, H. (2002). Flying and Handling qualities of a Fly-by-Wire Blended-Wing-Body civil


transport aircraft. PhD thesis, Cranfield University.

Denieul, Y., Alazard, D., Bordeneuve, J., Toussaint, C., and Taquin, G. (2015a). Interactions
of Aircraft Design and Control: Actuators Sizing and Optimization for an Unstable Blended
Wing-Body. In AIAA Atmospheric Flight Mechanics Conference, AIAA Aviation, Dallas, TX.
American Institute of Aeronautics and Astronautics.

Denieul, Y., Bordeneuve, J., Alazard, D., Toussaint, C., and Taquin, G. (2015b). Integrated
Design and Control of a Flying Wing Using Nonsmooth Optimization Techniques. In 3rd CEAS
EUROGNC Conference, Toulouse, France.

Doyle, J. C., Glover, K., Khargonekar, P. P., and Francis, B. A. (1989). State-space solutions to
standard H 2 and H∞ control problems. Automatic Control, IEEE Transactions on, 34(8):831–
847.

Drela, M. and Youngren, H. (2006). AVL Aerodynamic Analysis, Trim Calculation, Dynamic
Stability Analysis, Aircraft Configuration Development.

Durham, W. (1993). Constrained Control Allocation. 16(4):717–725.

Durham, W. (1997). Minimum Drag control allocation. JOURNAL OF GUIDANCE, CONTROL,


AND DYNAMICS, 20(1):190–193.

Durham, W. (2001). Computationnally efficient Control Allocation. JOURNAL OF GUIDANCE,


CONTROL, AND DYNAMICS, 24(3):519–524.

Durham, W. and Bordignon, K. (1995). Closed-Form Solutions to Constrained Control Allocation


Problem. JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS, 18(5):1000–1007.

Fathy, H., Reyer, J., Papalambros, P., and Ulsov, A. (2001). On the Coupling Between the Plant
and Controller Optimization Problems. In American Control Conference, 2001. Proceedings of
the, volume 3, pages 1864–1869 vol.3.

Favre, C. (1994). Fly-by-wire for commercial aircraft: the Airbus experience. International Journal
of Control, 59(1):139–157.
236 Bibliography

Feuersänger, A. P. (2007). Control of aircraft with reduced stability. PhD thesis, ISAE, Toulouse.

Fezans, N., Alazard, D., Imbert, N., and Carpentier, B. (2007). H∞ Control Design for Mul-
tivariable Mechanical Systems - Application to RLV Reentry. In AIAA Guidance, Navigation
and Control Conference and Exhibit, Guidance, Navigation, and Control and Co-located Con-
ferences. American Institute of Aeronautics and Astronautics.

Fezans, N., Alazard, D., Imbert, N., and Carpentier, B. (2008). H∞ control design for generalized
second order systems based on acceleration sensitivity function. Control and Automation, 2008
16th Mediterranean Conference on, pages 1508–1513.

FURUTA, K., KAJIWARA, H., and KOSUGE, K. (1980). Digital control of a double inverted
pendulum on an inclined rail. International Journal of Control, 32(5):907–924.

Gage, S. (2003). Creating a Unified Graphical Wind Turbulence Model from Multiple Specifi-
cations. In AIAA Modeling and Simulation Technologies Conference and Exhibit, Guidance,
Navigation, and Control and Co-located Conferences. American Institute of Aeronautics and
Astronautics.

Gahinet, P. and Apkarian, P. (1994). A linear matrix inequality approach to H∞ control. Inter-
national journal of robust and nonlinear control, 4(4):421–448.

Gahinet, P. and Apkarian, P. (2011a). Decentralized and fixed-structure H∞ control in MAT-


LAB. In Decision and Control and European Control Conference (CDC-ECC), 2011 50th IEEE
Conference on, pages 8205–8210. IEEE.

Gahinet, P. and Apkarian, P. (2011b). Structured H∞ synthesis in Matlab. In IFAC World


Congress, Università Cattolica del Sacro Cuore, Milano, Italy.

Gahinet, P. and Apkarian, P. (2012). Frequency-Domain Tuning of Fixed-Structure Control Sys-


tems. In UKACC International Conference on Control 2012, pages 178–183, Cardiff, UK. IEEE.

Gahinet, P. and Apkarian, P. (2013). Automated tuning of gain-scheduled control systems. In


Decision and Control (CDC), 2013 IEEE 52nd Annual Conference on, pages 2740–2745. IEEE.

Garmendia, D. C., Chakraborty, I., and Mavris, D. N. (2015a). Method for Evaluating Electrically
Actuated Hybrid Wing Body Control Surface Layouts. Journal of Aircraft, 52(6):1780–1790.

Garmendia, D. C., Chakraborty, I., and Mavris, D. N. (2015b). Multidisciplinary Approach to


Assessing Actuation Power of a Hybrid Wing Body. Journal of Aircraft, 53(4):900–913.

Garmendia, D. C., Chakraborty, I., Trawick, D. R., and Mavris, D. N. (2014). Assessment of
Electrically Actuated Redundant Control Surface Layouts for a Hybrid Wing Body Concept.
In 14th AIAA Aviation Technology, Integration, and Operations Conference, AIAA Aviation.
American Institute of Aeronautics and Astronautics.

Gaulocher, S., Roos, C., and Cumer, C. (2007). Aircraft Load Alleviation During Maneuvers
Using Optimal Control Surface Combinations. JOURNAL OF GUIDANCE, CONTROL, AND
DYNAMICS, 30(2):591–600.
Bibliography 237

Goldthorpe, S. H., Rossitto, K. F., Hyde, D. C., and Krothapalli (2010). X-48b Blended Wing
Body Flight Test Performance of Maximum Sideslip and High to Post Stall Angle-of-Attack
Command Tracking. In AIAA Guidance, Navigation, and Control Conference, pages 1–17,
Toronto, Ontario Canada. AIAA.

Hallberg, E. N. (1997). On integrated plant, control and guidance design. Technical report, DTIC
Document, Monterey CA.

Harkegard, O. (2004). Dynamic Control Allocation Using Constrained Quadratic Programming.


Journal of Guidance, Control, and Dynamics, 27(6):1028–1034.

Harkegard, O. and Glad, S. T. (2005). Resolving actuator redundancy- optimal control vs control
allocation. Automatica, 41(1):137–144.

Hartmann, G. L., Hauge, J. A., and Hendrick, R. C. (1976). F-8c digital CCV flight control laws.

Hovakimyan, N. and Cao, C. (2010). L1 adaptive control theory: guaranteed robustness with fast
adaptation, volume 21. Siam.

Jackson, E. B. and Buttrill, C. W. (2006). Control laws for a wind tunnel free-flight study of a
blended-wing-body aircraft. Technical report, NASA TM-2006-214501.

Johansen, T., Fossen, T., and Berge, S. (2004). Constrained Nonlinear Control Allocation With
Singularity Avoidance Using Sequential Quadratic Programming. IEEE TRANSACTIONS ON
CONTROL SYSTEMS TECHNOLOGY, 12(1):211–216.

Johansen, T. A. and Fossen, T. I. (2013). Control allocation- a survey. Automatica, 49(5):1087–


1103.

Kajita, S., Kanehiro, F., Kaneko, K., Yokoi, K., and Hirukawa, H. (2001). The 3d Linear Inverted
Pendulum Mode: A simple modeling for a biped walking pattern generation. volume 1, pages
239–246. IEEE.

Kalman, R. E. (1960). Contributions to the theory of optimal control. Bol. Soc. Mat. Mexicana,
5(2):102–119.

Kaminer, I. I., Howard, R. M., and Buttrill, C. S. (1997). Development of closed-loop tail-sizing
criteria for a High Speed Civil Transport. Journal of aircraft, 34(5):658–664.

Kozek, M. and Schirrer, A. (2014). Modeling and Control for a Blended Wing Body Aircraft.
Springer.

Lawrence, C., Zhou, J., and Tits, A. (2010). FSQP Manual.

Leman, T. J. (2009). L1 ADAPTIVE CONTROL AUGMENTATION SYSTEM FOR THE X-48B


AIRCRAFT. PhD thesis, University of Illinois.

Liao, F., Lum, K. Y., and Wang, J. L. (2005a). An LMI-based optimization approach for integrated
plant/output-feedback controller design. In American Control Conference, 2005. Proceedings
of the 2005, pages 4880–4885. IEEE.
238 Bibliography

Liao, F., Lum, K. Y., and Wang, J. L. (2005b). Mixed H2/H∞ Sub-Optimization Approach for
Integrated Aircraft/Controller Design. In presenting in the 16th IFAC World Congress, Prague,
Czech Republic.

Liebeck, R. (2005). Design of the Blended-Wing Body subsonic transport. 2005-06. von Karman
Institute for Fluid Dynamics.

Liebeck, R. (2015). Blended Wing Body X-48b Flight Test.

Liebeck, R., Page, M., and Rawdon, B. (1998). Blended-wing-body subsonic commercial transport.
In 36th AIAA Aerospace Sciences Meeting and Exhibit, Aerospace Sciences Meetings. American
Institute of Aeronautics and Astronautics.

Liebeck, R. H. (2004). Design of the Blended Wing Body Subsonic Transport. Journal of Aircraft,
41(1):10–25.

Mader, C. A. and Martins, J. R. R. A. (2013). Stability-Constrained Aerodynamic Shape Opti-


mization of Flying Wings. Journal of Aircraft, 50(5):1431–1449.

Martínez-Val, R., Cuerno, C., Pérez, E., and Ghigliazza, H. H. (2010). Potential Effects of Blended
Wing Bodies on the Air Transportation System. Journal of Aircraft, 47(5):1599–1604.

Martínez-Val, R. and Pérez, E. (2005). Medium Size Flying Wings. von Karman Institute for
Fluid Dynamics.

Martins, J. R. R. A. and Lambe, A. B. (2013). Multidisciplinary Design Optimization: A Survey


of Architectures. AIAA Journal, 51(9):2049–2075.

MATLAB (2014). version 2014b. Robust Control ToolboxTM The MathWorks Inc., Natick,
Massachusetts, USA.

Meheut, M., Arntz, A., and Carrier, G. (2012). Aerodynamic Shape Optimizations of a Blended
Wing Body Configuration for Several Wing Planforms. In AIAA, volume 10, page 3122, New
Orleans, Louisiana. American Institute of Aeronautics and Astronautics.

Niewoehner, R. J. and Kaminer, I. (1996). Integrated Aircraft-Controller Design using Linear


Matrix Inequalities. Journal of Guidance, Control, and Dynamics, 19(2):445–452.

Perez, R., Liu, H., and Behdinan, K. (2006). A Multidisciplinary Optimization Framework
for Control-Configuration Integration in Aircraft Conceptual Design. Journal of Aircraft,
43(6):1937–1948.

Piperni, P., DeBlois, A., and Henderson, R. (2013). Development of a Multilevel Multidisciplinary-
Optimization Capability for an Industrial Environment. AIAA Journal, 51(10):2335–2352.

Prigent, S., Belleville, M., Druot, T., Rondepierre, A., and Marechal, P. (2013). Chance con-
strained business case of a three-engines hybrid aircraft. In 10th World Congress on Structural
and Multidisciplinary Optimization, Orlando, florida, USA.

Puyou, G. (2005). Conception multi-objectifs de lois de pilotage pour un avion de transport civil.
PhD thesis, ISAE-Airbus, Toulouse, France.
Bibliography 239

Qin, N., Vavalle, A., Le Moigne, A., Laban, M., Hackett, K., and Weinerfelt, P. (2004). Aerody-
namic considerations of blended wing body aircraft. Progress in Aerospace Sciences, 40(6):321–
343.

Ravichandran, T., Wang, D., and Heppler, G. (2006). Simultaneous plant-controller design opti-
mization of a two-link planar manipulator. Mechatronics, 16(3–4):233–242.

Raymer, D. P. (1999). Aircraft Design: A Conceptual Approach. Inc., Reston, VA.

Ricardez, L. A. (2008). Simultaneous Design and Control of Chemical Plants: A Robust Modelling
Approach. PhD thesis, Waterloo, Canada.

Ricardez-Sandoval, L., Budman, H., and Douglas, P. (2009). Simultaneous design and control
of chemical processes with application to the Tennessee Eastman process. Journal of Process
Control, 19(8):1377–1391.

Roman, D., Allen, J., and Liebeck, R. (2000). Aerodynamic Design Challenges of the Blended-
Wing-Body Subsonic Transport. In 18th Applied Aerodynamics Conference. AIAA.

Roos, C. (2003). Roll Architecture Optimisation Method and Tool. Airbus Internship Report,
Airbus, Toulouse, France.

Roos, C., Hardier, G., and Biannic, J.-M. (2014). Polynomial and rational approximation with the
APRICOT Library of the SMAC toolbox. In Control Applications (CCA), 2014 IEEE Conference
on, pages 1473–1478.

Roskam, J. (1985a). Airplane design. DARcorporation.

Roskam, J. (1985b). Airplane Design Part IV: layout of landing gear and systems. DARcorporation,
Kansas.

Roskam, J. (1985c). Airplane Design Part VI: Preliminary Calculation of Aerodynamic, Thrust
and Power Characteristics. DARcorporation, Kansas.

Roskam, J. (1995). Airplane flight dynamics and automatic flight controls. DARcorporation.

Ruben Perez, Hugh Liu, and Kamran Behdinan (2004). Evaluation of Multidisciplinary Optimiza-
tion Approaches for Aircraft Conceptual Design. In 10th AIAA/ISSMO Multidisciplinary Analysis
and Optimization Conference, Multidisciplinary Analysis Optimization Conferences. American
Institute of Aeronautics and Astronautics.

Rynaski, E. G. and Weingarten, N. C. (1972). Flight Control Principles for Control-Configured Ve-
hicles. Technical Report Technical Report AFFDL-TR-71-154, Cornell Aeronautical Laboratory,
Inc, Air Force Flight Dynamics Laboratory Wright-Patterson Air Force Base, OHIO.

Sahasrabudhe, V., Celi, R., André, and Tits, L. (1997). Integrated Rotor-Flight Control System
Optimization with Aeroelastic and Handling Qualities Constraints. Journal of Guidance, Control,
and Dynamics, 20(2):217–224.
240 Bibliography

Sargeant, M. A., Hynes, T. P., Graham, W. R., Hileman, J. I., Drela, M., and Spakovszky, Z. S.
(2010). Stability of Hybrid-Wing-Body-Type Aircraft with Centerbody Leading-Edge Carving.
Journal of Aircraft, 47(3):970–974.

Saucez, M. (2013). Handling Qualities of the Airbus Flying Wing Resolution. PhD thesis, ISAE-
Airbus, Toulouse.

Saucez, M. and Boiffier, J.-L. (2012). Optimization of Engine Failure on a Flying Wing Configu-
ration. In AIAA Atmospheric Flight Mechanics Conference, Guidance, Navigation, and Control
and Co-located Conferences, Minneapolis, Minnesota. AIAA.

Silvestre, C., Pascoal, A., Kaminer, I., and Healey, A. (1998). Plant-Controller Optimization
with Applications to Integrated Surface Sizing and Feedback Controller Design for Autonomous
Underwater Vehicles (AUVs). In American Control Conference, volume 3, pages 1640–1644.

Sima, V. (1996). Algorithms for linear-quadratic optimization, volume 200. CRC Press.

Simon, E. and Wertz, V. (2011). Alternatives with stronger convergence than coordinate-descent
iterative LMI algorithms. arXiv preprint arXiv:1110.2615.

Stein, G. (2003). The practical, physical (and sometimes dangerous) consequences of control
must be respected, and the underlying principles must be clearly and well taught. IEEE Control
Systems Magazine, 272(1708/03).

Syrmos, V. L., Abdallah, C. T., Dorato, P., and Grigoriadis, K. (1997). Static output feedback—a
survey. Automatica, 33(2):125–137.

Taquin, G. (2009). Flight Mechanics- Master of Science.

Trélat, E. (2008). Contrôle optimal: théorie & applications. Vuibert.

Velni, J. M., Meisami-Azad, M., and Grigoriadis, K. M. (2009). Integrated damping parame-
ter and control design in structural systems for H2 and H∞ specifications. Structural and
Multidisciplinary Optimization, 38(4):377–387.

Walker, S. A. (1976). DESIGN OF A CONTROL CONFIGURED TANKER AIRCRAFT. NASA


Technical Report N76-31158, NASA, Wright-Patterson Air Force Base, Headquarters, Aeronau-
tical Systems Division (AFSC).

Waters, S. M., Voskuijl, M., Veldhuis, L. L., and Geuskens, F. J. (2013). Control allocation
performance for blended wing body aircraft and its impact on control surface design. Aerospace
Science and Technology, 29(1):18–27.

Wildschek, A. (2014). Flight Dynamics and Control Related Challenges for Design of a Commercial
Blended Wing Body Aircraft. In AIAA Guidance, Navigation, and Control Conference, National
Harbor, Maryland. AIAA.

Wise, K. A., Brinker, J. S., Calise, A. J., and Enns, D. F. (1999). Direct adaptive reconfig-
urable flight control for a tailless advanced fighter aircraft. International Journal of Robust and
Nonlinear Control, 9(14):999–1012.
Bibliography 241

Yang, G.-h. and Lum, K.-Y. (2003). An optimization approach to integrated aircraft-controller
design. In American Control Conference, 2003. Proceedings of the 2003, volume 2, pages
1649–1654. IEEE.

Zames, G. (1981). Feedback and optimal sensitivity: Model reference transformations, mul-
tiplicative seminorms, and approximate inverses. Automatic Control, IEEE Transactions on,
26(2):301–320.

Zhang, Y. and Lum, K.-Y. (2007). Integrated-optimal design of airplane and flight control using
genetic algorithms. Evolutionary Computation, 2007. CEC 2007. IEEE Congress on, pages
2980–2987.

Zhou, K., Doyle, J. C., and Glover, K. (1996). Robust and optimal control, volume 40. Prentice
Hall New Jersey.
Résumé — La prochaine génération d’avions civil sera probablement une révolution en terme
de configuration d’avion, différant largement de l’architecture désormais classique “fuselage- ailes
- moteurs sous voilure”.Du point de vue des qualités de vol, la tendance actuelle est d’évoluer vers
des avions de moins en moins stables, à la fois en longitudinal et latéral. Il est dès lors probable
que les futurs avions ne seront pas directement contrôlables par un humain sans l’apport de lois de
commande stabilisantes. Il devient alors nécessaire de considérer l’apport des systèmes de com-
mandes de vol très tôt dans la conception de l’avion, notamment pour le dimensionnement des
empennages, gouvernes et actionneurs, contrairement au processus actuel qui ne prend principale-
ment en compte que des critères “boucle ouverte” d’équilibre en phase de conception préliminaire.
Plutôt qu’un processus itératif de dimensionnement puis synthèse de lois de commande, nous
proposons d’optimiser simultanément les tailles de gouvernes, actionneurs et commandes de vol
en tenant compte des instabilités longitudinales et latérales, ainsi que des contraintes industrielles
sur la structure de correcteurs, sur un cas d’application de type aile volante. Ce processus de
“co-design” permet de dimensionner des paramètres physiques de l’avion en tenant compte des
apports d’une boucle de retour pour contrer des perturbations externes telles que de la turbulence
atmosphérique, permettant un avion plus sûr et optimal.

Mots clés : Dimensionnement avion, Systèmes commandes de vol, Lois de commande,


Optimisation, Qualités de vol.

Abstract — Next generation of civil transport aircraft is likely to be a radical change in overall
configuration compared to traditional tube-and-wing design. From a handling qualities perspective,
current trend in modern airliners is to evolve towards more and more unstable aircraft, both
from longitudinal and lateral-directional point of view. As a consequence future aircraft may not
be controllable by human operator without stabilizing control laws. It then becomes necessary
to consider flight control systems contribution early in the design phase for control surfaces,
empennages and actuators sizing, as opposed to traditional way of working dealing only with
open-loop criteria for preliminary sizing. Instead of an iterative process of sizing and control
laws synthesis, we propose to concurrently optimize control surfaces, actuators and flight control
laws taking into account longitudinal and lateral instability as well as industrial structure for
controllers, for unstable configurations such as Blended Wing-Body (BWB). This “co-design”
procedure enables sizing of physical aircraft parameters taking into account benefits from feedback
stabilization for counteracting external disturbance such as atmospheric turbulence, thus leading
to safer and more optimal aircraft configurations.

Keywords: Aircraft Design, Flight Control Systems, Control Laws, Optimization, Handling
Qualities.

ISAE-ONERA Commande des Systèmes et Dynamique du Vol- CSDV


2, av. Edouard Belin
FR-31055 TOULOUSE CEDEX 4

Vous aimerez peut-être aussi